Next Article in Journal
Characteristics of Conventional and Microwave Sintered Iron Ore Preform
Previous Article in Journal
Influence of Polyurea Coatings on Low-Longitudinal-Reinforcement-Ratio Reinforced Concrete Beams Subjected to Bending
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advance and Modification Strategies of Transition Metal Dichalcogenides (TMDs) in Aqueous Zinc Ion Batteries

1
School of Packaging and Materials Engineering, Hunan University of Technology, Zhuzhou 412007, China
2
School of Metallurgy and Environment, Central South University, Changsha 410083, China
*
Author to whom correspondence should be addressed.
Materials 2022, 15(7), 2654; https://doi.org/10.3390/ma15072654
Submission received: 10 March 2022 / Revised: 26 March 2022 / Accepted: 1 April 2022 / Published: 4 April 2022

Abstract

:
In recent years, aqueous zinc ion batteries (ZIBs) have attracted much attention due to their high safety, low cost, and environmental friendliness. Owing to the unique layered structure and more desirable layer spacing, transition metal dichalcogenide (TMD) materials are considered as the comparatively ideal cathode material of ZIBs which facilitate the intercalation/ deintercalation of hydrated Zn2+ between layers. However, some disadvantages limit their widespread application, such as low conductivity, low reversible capacity, and rapid capacity decline. In order to improve the electrochemical properties of TMDs, the corresponding modification methods for each TMDs material can be designed from the following modification strategies: defect engineering, intercalation engineering, hybrid engineering, phase engineering, and in-situ electrochemical oxidation. This paper summarizes the research progress of TMDs as cathode materials for ZIBs in recent years, discusses and compares the electrochemical properties of TMD materials, and classifies and introduces the modification methods of MoS2 and VS2. Meanwhile, the corresponding modification scheme is proposed to solve the problem of rapid capacity fading of WS2. Finally, the research prospect of other TMDs as cathodes for ZIBs is put forward.

Graphical Abstract

1. Introduction

Due to the deterioration of the environment and the deficit of fossil energy, it is increasingly important to develop environmentally friendly, sustainable, and renewable energy [1,2,3]. At present, some renewable energy power generation can meet the requirements of environmental protection and sustainable development, such as solar power, wind power, and tidal power generation, but these energy sources suffer from regional limitations and instability, limiting their wider application. As an important part of energy for sustainable development, electrochemical energy storage has become an active research field in recent decades [4,5,6]. In the current energy market, lithium-ion batteries (LIBs) are dominant in automobile, medical equipment, portable wearable equipment, and other industries because of their excellent energy density and good environmental performance [7,8]. However, most lithium-ion batteries use organic solvents as electrolytes, which may cause safety problems and increase costs [9,10,11,12]. Compared with non-aqueous batteries, aqueous rechargeable batteries have the characteristics of low cost, non-toxicity, and non-flammability, which makes them safer, more environmentally friendly, and more economical [13,14,15].
In addition to LIBs, many other rechargeable aqueous metal ion batteries have been developed, including sodium ion batteries [16], potassium ion batteries [17], aluminum ion batteries [18], calcium ion batteries [19], and zinc ion batteries. Compared with other rechargeable aqueous metal ion batteries, zinc ion batteries have many advantages [20,21,22]: (1) Zinc ion batteries can be directly assembled in the air without inert environment, which can reduce the battery assembly cost. (2) Zinc ions can be electrodeposited reversibly in aqueous solution, so the zinc sheet can be directly used as the anode of the battery. (3) Zinc as anode has a higher theoretical capacity and lower oxidation/reduction potential (–0.76 V) than the standard hydrogen electrode, which indicates that there is a higher open circuit voltage when coupled with cathode. Therefore, in many rechargeable water-based metal ion batteries, the research on aqueous zinc ion batteries is increasingly concerned.
As shown in Figure 1a, the aqueous zinc ion battery (ZIB) is mainly composed of the battery shell, cathode material, anode material, electrolyte, and diaphragm. There are many kinds of cathode materials, including manganese-based materials [23], vanadium-based materials [24], and Prussian blue analogues [25], all of which possess a certain capacity of Zn2+ storage. Since the theoretical interlayer spacing of transition metal disulfide compounds is larger than the diameter of Zn2+, Zn2+ can be intercalated and deintercalated in TMD material, which also indicates that TMD material is feasible as the cathode of ZIB [26]. At present, few studies have been conducted on TMD as a cathode material of ZIB, and there is not a systematic exposition, which also shows from the aspect that TMD as cathode of ZIB is quite novel. Figure 1b displays the trend of the increasing number of publications, evidencing the increasing attention paid to TMD materials in ZIB. The latest research progress of TMD series materials as cathodes of ZIB in recent years is reviewed in this paper, and several modification methods which can improve the Zn2+ storage capacity and structural stability of TMD materials are expounded.

2. Characterization and Synthetic Methods of TMD

Two-dimensional layered transition metal disulfides (TMDs) generally have X-M-X structure, where M is the transition metal, such as molybdenum, vanadium, tungsten, bismuth, and other metal elements, while X is generally sulfur, selenium, and other elements, as shown in Figure 2a. These layered structures facilitate the transport of various carriers and can also adapt to the volume change during ion insertion. Because of their different chemical composition and unique crystal structure, as well as the fact that the d-orbitals can be filled with different elements, TMDs can be used as functional materials for electronic insulation, semiconductors, and superconductivity [27].
Among all TMDs materials, molybdenum disulfide (MoS2) has received extensive attention as a typical representative. Molybdenum disulfide is composed of two layers of sulfur atoms and one layer of molybdenum atoms, with the metal molybdenum layer interposed between the two sulfur layers, alternately stacked to form a sandwich-like structure. The sulfur atoms are bound together by van der Waals force, while the S-Mo-S atoms are linked by strong covalent bonds [28,29]. MoS2 not only has a layered structure, but also has different phases (1T, 2H, and 3R), and each phase also has different physical properties and chemical characteristics [30]. As shown in Figure 2b, MoS2 in 2H and 3R phases both demonstrate the triangular prismatic coordination of Mo atoms, and 2H-MoS2 is very stable because of the two layers of units stacked in hexagon symmetry. Meanwhile, 3R-MoS2 has rhombus symmetry, and each unit has three layers. On the contrary, the Mo atoms of 1T-MoS2 (metal phase) are octahedral coordinated and most unstable [31,32]. VS2 is also a common material in TMDs, where the sulfur and vanadium layers are stacked together in a sandwich-like structure by van der Waals force interactions [33,34]. Due to the large interlayer spacing, VS2 has great potential in the intercalation/deintercalation of ions, such as Li+, Na+, Zn2+, Mg2+, and Al3+ [35,36,37].
Figure 2. (a) Illustration of TMD structure; (b) Atomic structures of 1T-, 2H-, and 3R-MoS2 [31].
Figure 2. (a) Illustration of TMD structure; (b) Atomic structures of 1T-, 2H-, and 3R-MoS2 [31].
Materials 15 02654 g002
In order to obtain layered and nanoflower-like sulfide materials, TMD materials are generally prepared by hydrothermal and solvothermal methods. The synthetic methods of TMD materials as ZIB cathode are summarized in Table 1.

3. TMDs as ZIBs Cathode

Layered MoS2, VS2, WS2, and VSe2 have been proven to be feasible cathode materials for ZIB in recent years. MoS2 is the earliest and most extensively material studied in ZIB owing to its unique layered structure and different phases. However, compared with manganese-based materials and vanadium-based materials, TMDs have lower conductivity and larger capacity changes, which will lead to the lower rate performance and poor cycle performance of electrode materials during charge and discharge, which is also the biggest challenge for TMDs materials. In order to improve the electrochemical performance of TMDs, different modification strategies of TMDs were designed. In this section, the research progress of MoS2, VS2, WS2, and VSe2 materials as cathode materials for ZIB is described, and various modification methods of MoS2 and VS2 materials are discussed and summarized.

3.1. MoS2 and Modification

Liu et al. [53] did not obtain obvious redox peaks in CV curves of MoS2 and WS2, indicating that it is difficult for MoS2 and WS2 to store zinc ions. The measured cyclic discharge capacities provided by MoS2 and WS2 were 18 mAh·g−1 and 22 mAh·g−1, respectively, and the EDS images of MoS2 and WS2 in the fully discharged state also had low zinc contents, so their conclusions showed that MoS2 and WS2 materials were not suitable as cathode materials for ZIB. The electrochemical performance of unmodified MoS2 in ZIB is undesirable, such that more and more scholars have paid attention to the modified MoS2 as the cathode material of ZIB.

3.1.1. Defect Engineering

Defect engineering is used to change the surface properties and structure of TMDs to improve the electrochemical performance, which is undoubtedly a preferable modification strategy. The possible vacancy defects in TMDs consist of sulfur vacancies, transition metal vacancies, edge vacancies, and holes in the crystal lattice [54,55,56]. The main effects of defects on electrode materials are divided into three types: (1) More storage and adsorption sites of foreign ions are provided to improve the battery capacity. The introduction of defects increases the number of active sites significantly, which improves the electrochemical performance. (2) It changes the local structure and charge distribution on the surface of electrode materials, which can improve the conductivity, cycle stability, and rate performance of electrode materials. (3) Defects can make the structure more flexible and stable when external ions are intercalated and deintercalated. Therefore, the introduction of defects in the crystal structure of TMDs and the selection of appropriate chemical modification molecules or groups using the defect sites constructed on the surface of TMDs can improve the electrical conductivity, reversible capacity, and initial Coulombic efficiency (ICE) and enhance the reversibility of Zn2+ intercalation/de intercalation, thus improving the electrochemical performance of TMD electrode materials in ZIB.
For example, Xu et al. [38] designed the first MoS2−x/Zn cell and confirmed that defect engineering could be used to enhance the Zn storage capacity of MoS2. They predicted that the edge vacancies and sulfur vacancies were the main adsorption sites of Zn2+ by DFT, while the original surface was inert in storing Zn2+ due to the lack of these vacancies. As shown in Figure 3, owing to the absence of some chalcogen atoms, there is a large number of vacancies existing in the neatly arranged sulfur layer. These vacancies include edge vacancies and sulfur vacancies, which can easily accommodate more Zn2+ for intercalation, resulting in a significant increase in the reversible capacity of ZIBs. Such a layered nanostructure with rich defects can also reduce the diffusion energy barrier and significantly increase the diffusion rate of Zn ions [57]. Figure 4a shows that the capacity of defect-free MoS2 nanosheets is very low at a current density of 100 mA·g−1, while the defect-rich MoS2 nanosheets provide a high reversible capacity of 135 mAh·g−1. After 1000 cycles at a current density of 1000 mA·g−1, the defect-rich MoS2 electrode still delivers a high reversible capacity of 88.6 mAh·g−1 with a capacity retention of 87.8%. The electrochemical reaction mechanism of the MoS2−x/Zn battery at cathode and anode can be expressed as:
Cathode:  MoS2−x + 0.36Zn2+ + 0.72e → Zn0.36MoS2−x
Anode:  Zn2+ + 2e → Zn

3.1.2. Interlayer Engineering

The non-bonded characteristics of 2D layered materials with unique vdW gaps between adjacent atomic layers allows the insertion of atoms, ions, and molecules without breaking covalent bonds through a chemical intercalation process, thus changing the physical and chemical properties of the material [58]. TMDs are generally composed of 2D single-molecule layers combined with van der Waals forces and weak chemical bonds stacked along the direction perpendicular to the layers, a structure that has a larger specific surface area and is also conducive to the insertion of small intercalators, such as ions and molecules. Figure 5 shows that the insertion of external ions can expand the interlayer spacing of TMDs.
Li et al. [39] prepared expanded interlayer distance E-MoS2 nanosheets vertically aligned on carbon fiber cloth by a one-step glucose-assisted hydrothermal method as cathode materials for ZIB. The layer spacing of commercial MoS2 is 0.62 nm, while the interlayer spacing of E-MoS2 nanosheets is 0.70 nm. The increased interlayer spacing is due to the incorporation of hydrated Na+ and NH4+ into the MoS2 crystal framework during the hydrothermal reaction. The E-MoS2 electrode also showed good performance with a specific capacity of 202.6 mAh·g−1 and a specific energy density of 148.2 Wh·kg−1 at 0.1 A·g−1 as well as a capacity retention rate of 98.6% after more than 600 cycles with good cycling stability. The electrochemical reaction mechanism of the cathode and anode of E-MoS2/Zn battery is expressed as:
Cathode:  xZn2+ + 2xe+ MoS2 → ZnxMoS2
Anode:  xZn2+ + 2xe → xZn
Liang et al. [40] made MoS2-O to store Zn ions more efficiently by adjusting the interlayer spacing and designing hydrophilicity tuning engineering. A small amount of oxygen (5%) was added to MoS2, and the incorporation of oxygen improved the hydrophilicity of MoS2-O and made it possible for water to intercalate in MoS2-O, which also increased the interlayer spacing of MoS2-O from 0.62 to 0.95 nm, as shown in Figure 4b. Furthermore, if a diffusing hydrated Zn2+ were to diffuse through a narrow void separation in MoS2 (S to S layer separation), i.e., 0.31 nm, instead of the 0.62 nm, i.e., Mo to Mo layer separation, interlayer engineering becomes even more crucial. As a result, the incorporation of oxygen increased the diffusion rate of Zn2+ by three orders of magnitude and the storage capacity of Zn2+ was increased by10 times. When the current density is 100 mA·g−1, the capacity reaches 232 mAh·g−1, which is higher than that of MoS2-based electrode materials in other ZIB.
Unusually, Zhang et al. [41] designed a material with crystalline water introduced into MoS2 nanoflowers as cathode materials for ZIB. The MoS2·nH2O nanoflowers obtained by the one-step hydrothermal method are composed of layered nanosheets with low stacking height and uniform distribution, as shown in Figure 4c. The insertion of crystalline water not only increases the interlayer distance of MoS2 to 0.65 nm, but also shields the electrostatic interaction between zinc ions and matrix materials, and lubricates and increases the diffusion ion channels as a way to promote the migration of Zn2+. Under the synergistic effect of crystal water lubrication and nanoflower structure, the reversible capacity of MoS2·nH2O was 165 mAh·g−1 at 0.1 A·g−1. At the high current density of 2 A·g−1, the capacity retention rate after 800 cycles was 88%.
Huang et al. [42] first designed a molybdenum disulfide/polyaniline (MoS2/PANI) hybrid material with overlapping heterostructures as an excellent cathode for ZIB. The MoS2/PANI prepared by the hydrothermal method is a micro-nanoflower assembled by some ultra-thin wrinkled nanosheets with a layer spacing of 1.03 nm, as shown in Figure 4d. An appropriate amount of PANI was inserted into the MoS2 matrix, thus expanding the interlayer spacing and shielding the electrostatic interaction between some Zn2+ and MoS2 substrates. The intercalation of polyaniline not only expands the Zn2+ diffusion channel, but also reduces the band gap between the conduction band and the valence band of MoS2, which improves the charge transfer efficiency. Compared with the original MoS2, the zinc storage performance of the MoS2/PANI hybrid cathode was significantly improved. At a current density of 1.0 A·g−1, it provides a high reversible capacity of 106.5 mAh·g−1 and maintains 86% capacity after 1000 cycles. This also provides a modification strategy for developing other TMD materials as good cathode materials for ZIB.

3.1.3. Hybridization

The conductivity of TMDs is poor, which limits the diffusion power of ions. In electrochemical energy storage, carbon nanotubes, carbon fiber cloth, and graphene are the most commonly used hybrid materials. Carbon nanotubes and graphene materials have unique structures, high electrical conductivity, good mechanical properties, and large specific surface areas [59,60]. As shown in Figure 6, coating with a certain material or hybridizing with carbon nanotubes, carbon cloth, and graphene on the surface of the TMD can increase the surface area and charge transfer ability of the material to improve the electrochemical performance of the composite.
Huang [43] dispersed multi-walled carbon nanotubes in aqueous solution mixed with glucose and deionized water, and prepared ultra-thin MoS2 nanosheets with expanded layer spacing directly grown on the surface of carbon nanotubes (CNTs) by a one-step hydrothermal method (as shown in Figure 4e) as the ZIB cathode material. When the interlayer spacing of MoS2 increases to 1.0 nm, the increase of interlayer spacing reduces the diffusion energy barrier of Zn2+ between layers, and the main chain of CNTs promotes the electron transfer of Zn2+ between layers. More importantly, the layered structure hinders the aggregation of MoS2 nanosheets, increases the surface area, and provides a large number of active sites for electrochemical reactions. At the current density of 0.1 A·g−1, the reversible capacity of the MoS2@CNTs electrode was about 180.0 mAh·g−1. When the current density increases to 3 A·g−1, it can maintain 51.7% capacity, and after 500 cycles (the current density is 1.0 A·g−1) it can maintain 80.1% capacity. Therefore, MoS2@CNTs as cathode can provide high reversible capacity and good cycle stability for ZIB.

3.1.4. Phase Engineering

Phase engineering has been proven to be an effective method to regulate the electrochemical performance of TMDs [61]. According to different stacking methods, the materials can be divided into three bulk phases: 1T, 2H and 3R. The cell of 1T phase is octahedral, and the cell of 2H phase is prismatic. Due to different atomic configurations, single-layer TMD materials can also exist in three different forms of 1T, 2H, and 3R, respectively. MoS2 is a typical TMD which has different phase structures, different metal coordination geometries, different stacking sequences, and different properties. For example, the natural bulk MoS2 crystal has a 2H phase with triangular prism coordination geometry, but the coordination of molybdenum and sulfur atoms is octahedral in 1T phase which shows metallicity [62,63]. In fact, the conductivity of 1T phase MoS2 is about 105 times higher than that of the 2H phase MoS2 corresponding to semiconductor [64,65]. It can be reasonably speculated that MoS2 with different phases exhibits different properties in application. Phase engineering has also been proven to be an effective strategy to regulate the catalytic activity of MoS2 [66].
For example, Liu et al. [44] prepared MoS2 with different phases by regulating the reaction temperature of the hydrothermal reaction, and most samples showed a flower-like structure of self-assembled nanosheets. As the synthesis temperature decreases, the interlayer spacing of MoS2 increases from 0.62 nm to 0.68 nm, as shown in Figure 4f. Different from MoS2 nanosheets corresponding to 2H, MoS2 nanosheets with high 1T phase content (about 70%) have excellent specific capacity, and the reason for this is that the Zn2+ diffusion barrier of 1T-MoS2 is much lower than that of 2H-MoS2. At a current density of 1.0 A·g−1, the reversible capacity of 1T-MoS2 is about 140 mAh·g−1, but the capacity decreases rapidly after 100 cycles. However, the capacity attenuation of 2H-MoS2 can be ignored after 400 cycles, showing good cycle performance. How to make MoS2 show excellent specific capacity and maintain stable cycle performance is still worth exploring.

3.2. VS2 and Modification

In recent years, vanadium disulfide (VS2) based nanomaterials have become potential electrode materials for various energy storage batteries due to their highly controllable structure and chemical composition. The VS2 layer shows metallicity and has a considerable electronic state at the Fermi level, which is conducive to promoting electron and electrochemical activity [67]. At present, using VS2 as the cathode of ZIB is still considered to be a relatively new method. So far, there are only a few reports on VS2 as the cathode of ZIB. For example, He et al. [45] firstly synthesized rose-like VS2 nanoflowers assembled by nanosheets with diameter of 5–8 μm and thickness of 50–100 nm, whose interlayer spacing is 0.576 nm, as shown in Figure 7a. These nanosheets can realize the intercalation/deintercalation of Zn2+ in VS2 nanosheets. At a current density of 0.05 A·g−1, VS2 nanosheets have a high capacity of 190.3 mAh·g−1 and an energy density of 123 Wh·kg−1. The capacity retention rate reached 98.0% after 200 cycles at a current density of 0.5 A·g−1, which confirmed the feasibility of VS2 as the cathode of ZIB. They summarize the electrochemical reactions occurring in the cathode of Zn/VS2 batteries, as shown in Equation (5):
VS2 + 0.09Zn2+ + 0.18e → Zn0.09VS2;
Zn0.09VS2 + 0.14Zn2+ + 0.28e → Zn0.23VS2
In other reports, VS2 was modified to obtain electrode materials with better electrochemical performance. Consistently, we reviewed the research progress of modified VS2 as electrode material for ZIB and summarized two corresponding modification strategies in this section.

3.2.1. Hybridization

Jiao et al. [46] synthesized hierarchical 1T-VS2 directly on stainless steel mesh (VS2@SS) by the hydrothermal method. The prepared VS2 was composed of bending nanosheets with a transverse size of 5–8 μm, forming a highly layered network flower structure from Figure 7b. There is a large number of interlayer channels for electrolyte penetration in VS2 nanoflowers, which enhances the contact area between electrolyte and VS2 and promotes the transmission of electrons and ions. Therefore, the VS2@SS electrode with VS2 loading of 4–5 mg·cm−2 has a high capacity of 198 mAh·g−1 at a current density of 50 mA·g−1. After 2000 cycles at a current density of 2 A·g−1, the capacity retention rate was about 80%, showing a superior cycle life. When the mass load of VS2 is about 11 mg·cm−2, the electrode can maintain 90% capacity in 600 cycles (only 0.017% loss per cycle), which indicates that the stability of the composite structure is enhanced after the hybrid of TMD material and SS.
Graphene and reduced graphene oxide (rGO) are idealized cathode materials with large specific surface area, excellent electrical conductivity, and excellent mechanical elasticity [68]. In addition, the end-functional group surface of rGO further provides the possibility of strong interface coupling between the active material and the conductive skeleton, and it also realizes the rapid charge transfer and high structural stability against repeated electrochemical cycles [69,70]. Due to the unique layered structure and large specific surface area of graphene and VS2, the electrochemical performance of the electrode can be further improved by hybridizing graphene and VS2. For example, Chen et al. [65] synthesized vertically grown ultra-thin VS2 nanosheets (rGO-VS2) on graphene sheets by the solvothermal method. The interlayer distance of these ultrathin VS2 nanosheets is about 0.97 nm, as shown in Figure 7c, which is much larger than that of commercial VS2 crystals (0.575 nm). They also calculated that the surface area of rGO-VS2 was 34.2 m2·g−1, which was larger than that of commercial VS2 (26.5 m2·g−1), indicating that the introduction of graphene increased the specific surface area of the composites. VS2 nanosheets are closely connected with graphene, which can effectively prevent the dissolution of VS2 and improve the cycle stability of the battery. Therefore, when the current density is 0.1 A·g−1, it has a high capacity of 238 mAh·g−1. At a high current density of 5 A g−1, the capacity is 190 mAh·g−1, and the capacity retention is over 93% after 1000 cycles.
Metal oxides, polymers, C3N4, and MXene can also be hybridized with TMD materials [71,72,73]. Pu et al. [48] prepared a rose-like VS2@VOOH material with hydrophilic VOOH coating for the first time. The size of VS2@VOOH is about 10 μm and a large number of nanosheets are assembled into uniform rose-like morphology, as shown in Figure 7d. Hydrophilic VOOH coating is conducive to the penetration of electrolyte in ZIB, and the structure is still intact even after O-H is exchanged with water in VOOH. Their research also showed that the O-H bond in VOOH could not only improve the permeation of electrolyte to the electrode, but also reduce the possibility of vanadium dissolution. Therefore, after 350 cycles at a current density of 1.5 A·g−1, the specific capacity of VS2@VOOH composite can reach 107.5 mAh·g−1. Even after 400 cycles at a current density of 2.5 A·g−1, the VS2@VOOH material still maintained a specific capacity of 91.4 mAh·g−1, which further verified the potential of the modified VS2 as a ZIB cathode.

3.2.2. In Suit Electrochemical Oxidation

Electrochemical oxidation, as a feasible and multi-purpose strategy, has been proven to enhance the electrochemical behavior of low-valent active substances relative to evolutionary forms and newly formed high-active substances [74,75,76,77]. As shown in Figure 8, electrochemical oxidation of the inserted ions into the corresponding oxides without changing the position of the ion insertion into the TMD can provide more active sites or enhance the hydrophilicity of the ions, which is beneficial to the diffusion of Zn2+ and the expansion of the TMD layer spacing. Yang et al. [49] prepared a hollow spherical VS2·NH3 material with layer spacing expansion as the cathode of ZIB. The VS2·NH3 hollow flower ball assembled by nanosheets has a porous structure, and the expanded layer spacing is 0.98 nm, as shown in Figure 7e. They found that during the first charge process, the expanded VS2·NH3 transformed into V2O5·nH2O with a layer spacing of 1.21 nm after electrochemical oxidation. Then, Zn2+ was intercalated/deintercalated in V2O5·nH2O with large interlayer spacing, which provided high capacity and good stability for subsequent charge–discharge cycles. When the current density is 0.1, 0.2, 0.5, 1.0, 2.0, and 5.0 A·g−1, the capacity of the Zn/VS2·NH3 battery is 390, 372, 320, 260, 189, and 107 mAh·g−1, respectively, which is much higher than the VS2-based cathode of other ZIB previously reported. It is worth noting that the electrode formed by in-situ electrochemical oxidation into V2O5·nH2O reached an amazing 110% capacity retention compared with the initial capacity after 2000 cycles at a high current density of 3 A·g−1.
Yu et al. [50] first used the in situ electrochemical pretreatment of VS2 in aqueous medium and obtained VS2/VOx heterostructures as cathode materials for ZIB. The fluffy VOx nanosheets are uniformly grown on VS2, forming an interwoven porous electrode rich in VS2 and VOx heterostructures. The VS2/VOx heterostructure combines the high conductivity of VS2 and the high chemical stability of VOx, which is conducive to regulating the intercalation of guest ions and improving the chemical stability of the VS2 skeleton. The step-by-step insertion of Zn2+ changed the buffer volume, making the reaction kinetics faster and more reversible. As shown in Figure 7f, the VS2/VOx electrode maintains 75% high capacity in 3000 cycles at a current density of 1 A·g−1. After 1000 cycles, VS2/VOx still maintained the original crystal structure and nanosheet morphology, which again proved that the VS2/VOx heterostructure had high reversibility. As a consequence, the design of TMD heterostructures by in situ electrochemical oxidation provides a new strategy for the efficient energy storage of ZIB.

3.3. WS2 and Modification

WS2 is a natural metal sulfide, where the S atom is located in the lattice position of the closely packed hexagonal structure. The plane where the W atom lies is sandwiched between two S layers and each W atom is coordinated with the S atom. The W layer and the S layer are stacked together through a weak van der Waals effect to form a prismatic structure [78,79]. The theoretical layer spacing of WS2 is 0.618 nm, while the diameter of Zn2+ is only 0.404~0.43 nm. Theoretically, it can also be used as the electrode material of the energy storage battery, but there are few reports on it as the ZIB cathode material. A report [53] on WS2 pointed out that the interlayer spacing of commercial WS2 was insufficient, and the inherent conductivity of Zn2+ for reversible intercalation between layers was low, which indicates that the storage capacity of commercial WS2 for Zn2+ was poor. Another reason for this is that 2H-WS2 has a large Zn2+ intercalation barrier, which seriously hinders the diffusion of Zn2+ to electrode materials.
Tang et al. [51] proved for the first time that 1T-WS2 nanosheets are a promising candidate cathode material for ZIB. Thus, 1T phase WS-200 is stacked by ultrathin nanosheets, and the interlayer spacing is expanded from 0.618 nm to about 0.90 nm. After the introduction of 1T-WS2, the capacity of an almost inactive Zn/WS2 battery increased by eight times, and the reversible capacity was 179.99 mAh·g−1 at a current density of 200 mA·g−1. However, WS2 has the problem of rapid capacity degradation in the long cycle at different current densities from Figure 9a. Their explanation for this phenomenon is that the layered WS2 structure collapsed and dissolved, and some 1T-WS2 dissolved in the electrolyte during the intercalation/deintercalation of Zn2+, which provides a research direction for improving the reversible capacity and cycle stability of WS2.
In order to improve the electrochemical performance of WS2, a modification strategy similar to MoS2 and VS2 can be designed to modify WS2, the most important of which is hybridization and intercalation. For example, Debela et al. [80] synthesized WS2@NGr material by hybridizing WS2 with nitrogen-doped graphite. The reversible capacity of WS2@NGr electrode reached 963 mAh·g−1. Ratha et al. [81] synthesized layered WS2 and reduced graphene oxide (RGO) hybrid materials by a simple hydrothermal method. WS2/RGO composites showed excellent cycle stability in supercapacitors. Liu et al. [82] successfully synthesized a highly stable 1T-WS2 nanobelt with a special zigzag chain superlattice structure. The NH4+ insertion not only extends the interlayer spacing of WS2 to 0.95 nm, but also makes the structure highly stable. Although the above three WS2 modified materials are not applied to ZIB, they also provide many strategies. Starting from the microstructure of WS2 and the storage mechanism of Zn2+ in WS2, the carbon-based materials and WS2 can be hybridized, or external ions, such as NH4+ and crystal water, can be used for intercalation, so as to overcome the challenge of rapid capacity decline faced by WS2.

3.4. Other Materials

VSe2 is also a typical layered structure, and the vanadium layer is alternately located between two selenium layers, forming a sandwich structure formed by van der Waals interaction [83,84]. Larger interlayer spacing (0.611 nm) provides sufficient transport channels and active sites for active cations. Moreover, the electronic coupling force between adjacent V4+ -V4+ induces metal properties and excellent electronic conductivity, making it an attractive candidate electrode material [85,86]. Wu et al. [52] prepared ultra-thin VSe2 nanosheets and used them as cathode materials for ZIB. As shown in Figure 9b, the specific capacities of VSe2 are 250.6 and 132.6 mAh·g−1 at 200 and 5000 mA·g−1, respectively. After 800 cycles at 2 A·g−1, the capacity retention was 83%. The factors influencing the excellent electrochemical performance of VSe2 nanosheets cathode are attributed as follows: (1) Highly reversible Zn2+ intercalation/deintercalation storage mechanism; (2) Thanks to its ultra-thin two-dimensional morphology, Zn2+ has rapid diffusion kinetics (DZn2+ ≈ 10−8 cm−2 s−1); (3) The metal characteristics of VSe2 are conducive to the thermodynamics and kinetics of Zn2+ storage process; (4) Stability of crystal structure of VSe2 electrode material during long cycle. They summarized the storage mechanism of VSe2/Zn battery as 6 and 7:
Cathode:  VSe2 + 0.23Zn2+ + 0.46e → Zn0.23VSe2
Zn0.23VSe2 + 0.17Zn2+ + 0.34e → Zn0.4VSe2
Anode:  0.4Zn2+ + 0.8e → 0.4Zn
A few other TMD materials can also be used in the water-based zinc ion battery. For example, Zhang et al. [87] proved that CoS2 could be used as anode material for a sulfur redox type non-aqueous zinc battery. Li et al. [88] synthesized a novel pre-cured titanium disulfide (Na0.14TiS2) as anode material for ZIB. However, these materials are all anode materials for ZIB, which are not described too much here.

4. Conclusions and Future Work

Since ZIB offer good safety and environmental protection, and with the deepening research on ZIB, they are expected to become a substitute for LIB. The special layered structure of TMDs is conducive to the transport of carriers such as Zn2+. The nanoflower-like structure synthesized by the ordered stacking of monolayers confers TMD materials with a larger specific surface area. Some TMDs also have interlayer spacing larger than the diameter of hydrated zinc ion. Therefore, TMD materials are considered to represent a candidate cathode of ZIB with significant potential. Although TMDs are conductive, they also face the challenges of low reversible capacity and rapid capacity decay.
In recent years, some studies have reported that the modified TMD materials have excellent reversible capacity and cycle stability as cathodes of ZIB. Based on these research reports, the research progress of typical TMDs materials, such as MoS2, VS2, WS2, and VSe2, as cathodes of ZIB was described in this paper. The reversible capacity and long-period cycle performance are summarized in Table 2 and compared. At the same time, the modification ideas of TMD cathode materials for ZIB were combed, and the corresponding modification methods were summarized and discussed: (1) defect engineering, providing more ion storage sites and active sites by constructing chemical defects; (2) interlayer engineering, namely through the intercalation of guest ions to expand the layer spacing; (3) hybridization with carbon-based materials or other materials, making the microstructure of the composites more stable or maintained in a stable phase structure; (4) phase engineering, controlling the mutual transformation between 1H and 2R phases to reduce the diffusion energy barrier of Zn2+ and make the structure more stable; (5) in-situ electrochemical oxidation, without changing the position of ions in the interlayer, where the low-valent active substances are oxidized into kinetic high-valent active substances by electrochemistry.
Although these modification strategies are currently only applied to MoS2 and VS2, they should also be applicable to other TMD materials because the TMDs have similar special layered structures. For example, in view of the rapid decline of WS2 capacity, we may venture to assume that the hybridization of WS2 and carbon-based materials improves the structural stability of the composites, thus improving the poor cycle stability of WS2 electrodes. With the modification strategies put forward, TMD materials are bound to receive more attention, and more series members of TMDs will become excellent cathode materials for ZIB.

Author Contributions

Conceptualization, T.L. and Y.X.; software, T.L. and J.Y.; validation, H.L., Y.X., and A.S.; investigation, T.L. and H.L.; resources, A.S. and Y.L.; data curation, T.L.; writing—original draft preparation, T.L.; writing—review and editing, A.S.; supervision, A.S.; funding acquisition, A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Hunan Provincial Natural Science Foundation (2021JJ30215), the Scientific Research Project of Hunan Provincial Department of Education (21A0363) and the Education Reform Project in Colleges and Universities of Hunan Province (HNJG-2020-0572).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yan, H.; Zhang, X.; Yang, Z.; Xia, M.; Xu, C.; Liu, Y.; Yu, H.; Zhang, L.; Shu, J. Insight into the electrolyte strategies for aqueous zinc ion batteries. Coord. Chem. Rev. 2021, 452, 214297. [Google Scholar] [CrossRef]
  2. Yang, S.; Cheng, Y.; Xiao, X.; Pang, H. Development and application of carbon fiber in batteries. Chem. Eng. J. 2020, 384, 123294. [Google Scholar] [CrossRef]
  3. Wang, J.; Yang, Y.; Zhang, Y.; Li, Y.; Sun, R.; Wang, Z.; Wang, H. Strategies towards the challenges of zinc metal anode in rechargeable aqueous zinc ion batteries. Energy Storage Mater. 2020, 35, 19. [Google Scholar] [CrossRef]
  4. Zhou, H.; Li, X.; Li, Y.; Zheng, M.; Pang, H. Applications of MxSey (M = Fe, Co, Ni) and their composites in electrochemical energy storage and conversion. Nano-Micro Lett. 2019, 11, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Xiao, X.; Zou, L.; Pang, H.; Xu, Q. Synthesis of micro/nanoscaled metal–organic frameworks and their direct electrochemical applications. Chem. Soc. Rev. 2020, 49, 301. [Google Scholar] [CrossRef]
  6. Hua, Y.; Li, X.; Chen, C.; Pang, H. Cobalt Based metal-organic frameworks and their derivatives for electrochemical energy conversion and storage. Chem. Eng. J. 2019, 370, 37. [Google Scholar] [CrossRef]
  7. Zhang, N.; Xiao, X.; Pang, H. Transition metal (Fe, Co, Ni) fluoride-based materials for electrochemical energy storage. Nanoscale Horiz. 2019, 4, 99. [Google Scholar] [CrossRef]
  8. Wang, F.; Liu, Y.; Zhao, Y.; Wang, Y.; Wang, Z.; Zhang, W.; Ren, F. Facile synthesis of two-dimensional porous MgCo2O4 nanosheets as anode for lithium-ion batteries. Appl. Sci. 2017, 8, 22. [Google Scholar] [CrossRef] [Green Version]
  9. Qi, S.; Wu, D.; Dong, Y.; Liao, J.; Foster, C.W.; O’Dwyer, C.; Feng, Y.; Liu, C.; Ma, J. Cobalt-based electrode materials for sodium-ion batteries. Chem. Eng. J. 2019, 370, 185. [Google Scholar] [CrossRef]
  10. Zheng, M.; Chi, Y.; Hu, Q.; Tang, H.; Jiang, X.; Zhang, L.; Zhang, S.; Pang, H.; Xu, Q. Carbon nanotube-based materials for lithium–sulfur batteries. J. Mater. Chem. A 2019, 7, 17204. [Google Scholar] [CrossRef]
  11. Sharma, L.; Gond, R.; Senthilkumar, B.; Roy, A.; Barpanda, P. Fluorophosphates as efficient bifunctional electrocatalysts for metal air batteries. ACS Catal. 2020, 10, 43. [Google Scholar] [CrossRef]
  12. Ru, Y.; Zheng, S.; Xue, H.; Pang, H. Potassium cobalt hexacyanoferrate nanocubic assemblies for high performance aqueous aluminum ion batteries. Chem. Eng. J. 2020, 382, 122853. [Google Scholar] [CrossRef]
  13. Wrogemann, J.M.; Künne, S.; Heckmann, A.; Rodríguez-Pérez, I.A.; Siozios, V.; Yan, B.; Li, J.; Winter, M.; Beltrop, K.; Placke, T. Development of safe and sustainable dual ion batteries through hybrid aqueous/nonaqueous electrolytes. Adv. Energy Mater. 2020, 10, 1902709. [Google Scholar] [CrossRef] [Green Version]
  14. Zeng, X.; Hao, J.; Wang, Z.; Mao, J.; Guo, Z. Recent progress and perspectives on aqueous zn-based rechargeable batteries with mild aqueous electrolytes. Energy Storage Mater. 2019, 20, 410. [Google Scholar] [CrossRef]
  15. Kasiri, G.; Glenneberg, J.; Hashemi, A.B.; Kun, R.; Mantia, F.L. Mixed copper-zinc hexacya-noferrates as cathode materials for aqueous zinc-ion batteries. Energy Storage Mater. 2019, 19, 360. [Google Scholar] [CrossRef]
  16. Jin, T.; Ji, X.; Wang, P.; Zhu, K.; Zhang, J.; Cao, L.; Chen, L.; Cui, C.; Deng, T.; Liu, S.; et al. High energy aqueous sodium-ion batteries. Angew. Chem. Int. Ed. 2021, 60, 11943–11948. [Google Scholar] [CrossRef]
  17. Zhu, K.; Li, Z.; Jin, T.; Jiao, L. Low defects potassium cobalt hexacyanoferrate as a superior cathode for aqueous potassium ion batteries. J. Mater. Chem. A. 2020, 8, 21103–21109. [Google Scholar] [CrossRef]
  18. Pang, Q.; Yang, S.; Yu, X.; He, W.; Zhang, S.; Tian, Y.; Xing, M.; Fu, Y.; Luo, X. Realizing reversible storage of trivalent aluminum ions using VOPO4·2H2O nanosheets as cathode material in aqueous aluminum metal batteries. J. Alloy. Compd. 2021, 885, 161008. [Google Scholar] [CrossRef]
  19. Shi, Z.; Wu, J.; Ni, M.; Guo, Q.; Zan, F.; Xia, H. Superior performance of calcium birnessite by electrochemical conversion as cathode for aqueous calcium ion battery. Mater. Res. Bull. 2021, 144, 111475. [Google Scholar] [CrossRef]
  20. Fang, G.; Zhou, J.; Pan, A.; Liang, S. Recent advances in aqueous zinc ion batteries. ACS Energy Lett. 2018, 3, 2480–2501. [Google Scholar] [CrossRef]
  21. Konarov, A.; Voronina, N.; Jo, J.H.; Bakenov, Z.; Sun, Y.; Myung, S. Present and future perspective on electrode materials for rechargeable zinc-ion batteries. ACS Energy Lett. 2018, 3, 2620–2640. [Google Scholar] [CrossRef]
  22. Ming, J.; Guo, J.; Xia, C.; Wang, W.; Alshareef, H.N. Zinc-ion batteries: Materials, mechanisms, and applications. Mater. Sci. Eng. R Rep. 2019, 135, 58–84. [Google Scholar] [CrossRef]
  23. Zhang, T.; Tang, Y.; Fang, G.; Zhang, C.; Zhang, H.; Guo, X.; Cao, X.; Zhou, J.; Pan, A.; Liang, S. Electrochemical activation of manganese based cathode in aqueous Zinc-Ion electrolyte. Adv. Funct. Mater. 2020, 30, 2002711. [Google Scholar] [CrossRef]
  24. Ding, J.; Gao, H.; Ji, D.; Zhao, K.; Wang, S.; Cheng, F. Vanadium-based cathodes for aqueous zinc-ion batteries: From crystal structures, diffusion channels to storage mechanisms. J. Mater. Chem. A 2021, 9, 5258–5275. [Google Scholar] [CrossRef]
  25. Cao, T.; Zhang, F.; Chen, M.; Shao, T.; Li, Z.; Xu, Q.; Cheng, D.; Liu, H.; Xia, Y. Cubic manganese potassium hexacyanoferrate regulated by controlling of the water and fefects as a high capacity and stable cathode material for rechargeable aqueous zinc-ion batteries. ACS Appl. Mater. Interfaces 2021, 13, 26924–26935. [Google Scholar] [CrossRef] [PubMed]
  26. Lee, W.S.V.; Xiong, T.; Wang, X.; Xue, J. Unraveling MoS2 and transition metal dichalcogenides as functional zinc-ion battery cathode: A perspective. Small Methods 2020, 5, 2000815. [Google Scholar] [CrossRef] [PubMed]
  27. Wu, J.; Ciucci, F.; Kim, J.-K. Molybdenum disulfide based nanomaterials for rechargeable batteries. Chem. Eur. J. 2020, 26, 6296–6319. [Google Scholar] [CrossRef] [PubMed]
  28. Zhao, X.; Sui, J.; Li, F.; Fang, H.; Wang, H.; Li, J.; Cai, W.; Cao, G. Lamellar MoSe2 nanosheets embedded with MoO2 nanoparticles: Novel hybrid nanostructures promoted excellent performances for lithium ion batteries. Nanoscale 2016, 8, 17902–17910. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, L.; Xu, Z.; Wang, W.; Bai, X. Atomic mechanism of dynamic electrochemical lithiation processes of MoS2 nanosheets. J. Am. Chem. Soc. 2014, 136, 6693–6697. [Google Scholar] [CrossRef]
  30. Zhang, G.; Liu, H.; Qu, J.; Li, J. Two-dimensional layered MoS2: Rational design, properties and electrochemical applications. Energy Environ. Sci. 2016, 9, 1190. [Google Scholar] [CrossRef]
  31. Kuc, A.; Heine, T. The electronic structure calculations of two-dimensional transition-metal dichalcogenides in the presence of external electric and magnetic fields. Chem. Soc. Rev. 2015, 44, 2603. [Google Scholar] [CrossRef] [PubMed]
  32. Jiao, Y.; Hafez, A.M.; Cao, D.; Mukhopadhyay, A.; Ma, Y.; Zhu, H. Metallic MoS2 for high performance energy storage and energy conversion. Small 2018, 14, 1800640. [Google Scholar] [CrossRef]
  33. Wu, D.; Wang, C.; Wu, M.; Chao, Y.; He, P.; Ma, J. Porous bowl-shaped VS2 nanosheets/graphene composite for high rate lithium-ion storage. J. Energy Chem. 2020, 43, 24–32. [Google Scholar] [CrossRef] [Green Version]
  34. Jing, P.; Lu, H.; Yang, W.; Cao, Y. Interlayer-expanded and binder-free VS2 nanosheets assemblies for enhanced Mg2+ and Li+/Mg2+ hybrid ion storage. Electrochim. Acta 2020, 330, 135263–135272. [Google Scholar] [CrossRef]
  35. Zhu, J.; Jian, T.; Wu, Y.; Ma, W.; Lu, Y.; Sun, L.; Meng, F.; Wang, B.; Cai, F.; Gao, J.; et al. A highly stable aqueous Zn/VS2 battery based on an intercalation reaction. Appl. Surf. Sci. 2021, 544, 148882. [Google Scholar] [CrossRef]
  36. Kim, H.J.; Choi, B.K.; Lee, I.H.; Kim, M.J.; Chun, S.; Jozwiak, C.; Bostwick, A.; Rotenberg, E.; Chang, Y.J. Electronic structure and charge-density wave transition in monolayer VS2. Curr. Appl. Phys. 2021, 30, 8–13. [Google Scholar] [CrossRef]
  37. Liu, J.; Peng, W.; Li, Y.; Zhang, F.; Fan, X. A VS2@N-doped carbon hybrid with strong interfacial interaction for high-performance rechargeable aqueous Zn-ion batteries. J. Mater. Chem. C 2021, 9, 6308–6315. [Google Scholar] [CrossRef]
  38. Xu, W.; Sun, C.; Zhao, K.; Cheng, X.; Rawal, S.; Xu, Y.; Wang, Y. Defect engineering activating (boosting) zinc storage capacity of MoS2. Energy Storage Mater. 2019, 16, 527–534. [Google Scholar] [CrossRef]
  39. Li, H.; Yang, Q.; Mo, F.; Liang, G.; Liu, Z.; Tang, Z.; Ma, L.; Liu, J.; Shi, Z.; Zhi, C. MoS2 nanosheets with expanded interlayer spacing for rechargeable aqueous Zn-Ion batteries. Energy Storage Mater. 2019, 19, 94–101. [Google Scholar] [CrossRef]
  40. Liang, H.; Cao, Z.; Ming, F.; Zhang, W.; Anjum, D.H.; Cui, Y.; Cavallo, L.; Alshareef, H.N. Aqueous Zinc-Ion storage in MoS2 by tuning the intercalation energy. Nano Lett. 2019, 19, 3199. [Google Scholar] [CrossRef]
  41. Zhang, Z.; Li, W.; Wang, R.; Li, H.; Yan, J.; Jin, Q.; Feng, P.; Wang, K.; Jiang, K. Crystal water assisting MoS2 nanoflowers for reversible zinc storage. J. Alloy. Compd. 2021, 872, 159599. [Google Scholar] [CrossRef]
  42. Huang, M.; Mai, Y.; Zhao, L.; Liang, X.; Fang, Z.; Jie, X. Tuning the kinetics of zinc ion in MoS2 by polyaniline intercalation. Electrochim. Acta 2021, 388, 138624. [Google Scholar] [CrossRef]
  43. Huang, M.; Mai, Y.; Zhao, L.; Liang, X.; Fang, Z.; Jie, X. Hierarchical MoS2@CNTs hybrid as a long life and high rate cathode for aqueous rechargeable Zn-Ion batteries. ChemElectroChem 2020, 7, 4218–4223. [Google Scholar] [CrossRef]
  44. Liu, J.; Xu, P.; Liang, J.; Liu, H.; Peng, W.; Li, Y.; Zhang, F.; Fan, X. Boosting aqueous zinc-ion storage in MoS2 via controllable phase. Chem. Eng. J. 2020, 389, 124405. [Google Scholar] [CrossRef]
  45. He, P.; Yan, M.; Zhang, G.; Sun, R.; Chen, L.; An, Q.; Mai, L. Layered VS2 nanosheet based aqueous zn ion battery cathode. Adv. Energy Mater. 2017, 7, 1601920. [Google Scholar] [CrossRef]
  46. Jiao, T.; Yang, Q.; Wu, S.; Wang, Z.; Chen, D.; Shen, D.; Liu, B.; Cheng, J.; Li, H.; Ma, L.; et al. Binder-free hierarchical VS2 electrodes for high-performance aqueous zn ion batteries towards commercial level mass loading. J. Mater. Chem. A 2019, 7, 16330. [Google Scholar] [CrossRef]
  47. Chen, T.; Chen, X.; Zhang, Q.; Li, Y.; Peng, W.; Zhang, F.; Fan, X. VS2 nanosheets vertically grown on graphene as high-performance cathodes for aqueous zinc-ion batteries. J. Power Sources 2021, 477, 228652. [Google Scholar] [CrossRef]
  48. Pu, X.; Song, T.; Tang, L.; Tao, Y.; Cao, T.; Xu, Q.; Liu, H.; Wang, Y.; Xia, Y. Rose-like vanadium disulfide coated by hydrophilic hydroxyvanadium oxide with improved electrochemical performance as cathode material for aqueous zinc-ion batteries. J. Power Sources 2019, 437, 226917. [Google Scholar] [CrossRef]
  49. Yang, M.; Wang, Z.; Ben, H.; Zhao, M.; Luo, J.; Chen, D.; Lu, Z.; Wang, L.; Liu, C. Boosting the zinc ion storage capacity and cycling stability of interlayer-expanded vanadium disulfide through in-situ electrochemical oxidation strategy. J. Colloid Interface Sci. 2022, 607, 68–75. [Google Scholar] [CrossRef] [PubMed]
  50. Yu, D.; Wei, Z.; Zhang, X.; Zeng, Y.; Wang, C.; Chen, G.; Shen, Z.X.; Du, F. Boosting Zn2+ and NH4+ storage in aqueous media via in-situ electrochemical induced VS2/VOx heterostructures. Adv. Funct. Mater. 2021, 31, 2008743. [Google Scholar] [CrossRef]
  51. Tang, B.; Tian, N.; Jiang, J.; Li, Y.; Yang, J.; Zhu, Q. Investigation of zinc storage capacity of WS2 nanosheets for rechargeable aqueous zn-ion batteries. J. Alloy. Compd. 2022, 894, 162391. [Google Scholar] [CrossRef]
  52. Wu, Z.; Lu, C.; Wang, Y.; Zhang, L.; Jiang, L.; Tian, W.; Cai, C.; Gu, Q.; Sun, Z.; Hu, L. Ultrathin VSe2 nanosheets with fast ion diffusion and robust structural stability for rechargeable zinc-ion battery cathode. Small 2020, 16, 2000698. [Google Scholar] [CrossRef] [PubMed]
  53. Liu, W.; Hao, J.; Xu, C.; Mo, J.; Dong, L.; Jiang, F.; Wu, J.; Kang, Z.; Jiang, B.; Kang, F. Investigation of zinc ion storage of transition metal oxides, sulfides, and borides in zinc ion battery systems. Chem. Commun. 2017, 53, 6872–6874. [Google Scholar] [CrossRef]
  54. Chee, S.S.; Lee, W.J.; Jo, Y.R.; Cho, M.K.; Chun, D.; Baik, H.; Kim, B.J.; Yoon, M.H.; Lee, K.; Ham, M.H. Atomic vacancy control and elemental substitution in a monolayer molybdenum disulfide for high performance optoelectronic device arrays. Adv. Funct. Mater. 2020, 30, 1908147. [Google Scholar] [CrossRef]
  55. Shu, H.; Zhou, D.; Li, F.; Cao, D.; Chen, X. Defect engineering in MoSe2 for the hydrogen evolution reaction: From point defects to edges. ACS Appl. Mater. Interfaces 2017, 9, 42688–42698. [Google Scholar] [CrossRef]
  56. Zhang, Y.; Tao, L.; Xie, C.; Wang, D.; Zou, Y.; Chen, R.; Wang, Y.; Jia, C.; Wang, S. Defect engineering on electrode materials for rechargeable batteries. Adv. Mater. 2020, 32, 1905923. [Google Scholar] [CrossRef]
  57. Shao, Y.; Fang, Y.; Li, T.; Wang, Q.; Dong, Q.; Deng, Y.; Yuan, Y.; Wei, H.; Wang, M.; Gruverman, A.; et al. Grain boundary dominated ion migration in polycrystalline organic–inorganic halide perovskite films. Energy Environ. Sci. 2016, 9, 1752–1759. [Google Scholar] [CrossRef]
  58. Zhou, J.; Lin, Z.; Ren, H.; Duan, X.; Shakir, I.; Huang, Y.; Duan, X. Layered intercalation materials. Adv. Mater. 2021, 33, 2004557. [Google Scholar] [CrossRef]
  59. Cataldo, S.; Salice, P.; Menna, E.; Pignataro, B. Carbon nanotubes and organic solar cells. Energy Environ. Sci. 2012, 5, 5919–5940. [Google Scholar] [CrossRef]
  60. Wang, Y.; Shi, Z.; Huang, Y.; Ma, Y.; Wang, C.; Chen, M.; Chen, Y. Supercapacitor devices based on graphene materials. J. Phys. Chem. C. 2009, 113, 13103–13107. [Google Scholar] [CrossRef]
  61. Leng, K.; Chen, Z.; Zhao, X.; Tang, W.; Tian, B.; Nai, C.T.; Zhou, W.; Loh, K.P. Phase restructuring in transition metal dichalcogenides for highly stable energy storage. ACS Nano 2016, 10, 9208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Fan, X.; Xu, P.; Li, Y.C.; Zhou, D.; Sun, Y.; Nguyen, M.A.; Terrones, M.; Mallouk, T.E. Controlled exfoliation of MoS2 crystals into trilayer nanosheets. J. Am. Chem. Soc. 2016, 138, 5143–5149. [Google Scholar] [CrossRef] [PubMed]
  63. Tang, Q.; Jiang, D. Mechanism of hydrogen evolution reaction on 1T-MoS2 from first principles. ACS Catal. 2016, 6, 4953–4961. [Google Scholar] [CrossRef]
  64. Lei, Z.; Zhan, J.; Tang, L.; Zhang, Y.; Wang, Y. Recent development of metallic (1T) phase of Molybdenum disulfide for energy conversion and storage. Adv. Energy Mater. 2018, 8, 1703482. [Google Scholar] [CrossRef]
  65. Wang, R.; Yu, Y.; Zhou, S.; Li, H.; Wong, H.; Luo, Z.; Gan, L.; Zhai, T. Strategies on phase control in transition metal dichalcogenides. Adv. Funct. Mater. 2018, 28, 1802473. [Google Scholar] [CrossRef]
  66. Tan, C.; Luo, Z.; Chaturvedi, A.; Cai, Y.; Du, Y.; Gong, Y.; Huang, Y.; Lai, Z.; Zhang, X.; Zheng, L.; et al. Preparation of high percentage 1T-Phase transition metal dichalcogenide nanodots for electrochemical hydrogen evolution. Adv. Mater. 2018, 30, 1705509. [Google Scholar] [CrossRef]
  67. Jing, Y.; Zhou, Z.; Cabrera, C.R.; Chen, Z. Metallic VS2 Monolayer: A promising 2D anode material for lithium ion batteries. J. Phys. Chem. C 2013, 117, 25409. [Google Scholar] [CrossRef]
  68. Raccichini, R.; Varzi, A.; Passerini, S.; Scrosati, B. The role of graphene for electrochemical energy storage. Nat. Mater. 2015, 14, 271–279. [Google Scholar] [CrossRef] [PubMed]
  69. Keyu, X.; Kai, Y.; Xin, L.; Wei, L.; Chao, S.; Chenglu, L.; Robert, V.; Pulickel, A.; Bingqing, W. Superior potassium ion storage via vertical MoS2 “Nano-Rose” with expanded interlayers on graphene. Small 2017, 13, 1701471. [Google Scholar]
  70. Xiang, T.; Fang, Q.; Xie, H.; Wu, C.; Wang, C.; Zhou, Y.; Liu, D.; Chen, S.; Khalil, A.; Tao, S.; et al. Vertical 1T-MoS2 nanosheets with expanded interlayer spacing edged on a graphene frame for high rate lithium-ion batteries. Nanoscale 2017, 9, 6975–6983. [Google Scholar] [CrossRef]
  71. Liu, H.; Zhang, F.; Li, W.; Zhang, X.; Lee, C.-S.; Wang, W.; Tang, Y. Porous tremella-like MoS2/polyaniline hybrid composite with enhanced performance for lithium-ion battery anodes. Electrochim. Acta 2015, 167, 132. [Google Scholar] [CrossRef]
  72. Prabhu, P.; Jose, V.; Lee, J.M. Design strategies for development of TMD-Based heterostructures in electrochemical energy systems. Matter 2020, 2, 526. [Google Scholar] [CrossRef] [Green Version]
  73. Chen, Y.; Song, B.; Tang, X.; Lu, L.; Xue, J. Ultrasmall Fe3O4 nanoparticle/MoS2 nanosheet composites with superior performances for lithium ion batteries. Small 2014, 10, 1536. [Google Scholar] [CrossRef] [PubMed]
  74. Luo, H.; Wang, B.; Wang, F.; Yang, J.; Wu, F.; Ning, Y.; Zhou, Y.; Wang, D.; Liu, H.; Dou, S. Anodic oxidation strategy toward structure-optimized V2O3 cathode via electrolyte regulation for Zn-Ion storage. ACS Nano 2020, 14, 7328–7337. [Google Scholar] [CrossRef] [PubMed]
  75. Ding, J.; Gao, H.; Zhao, K.; Zheng, H.; Zhang, H.; Han, L.; Wang, S.; Wu, S.; Fang, S.; Cheng, F. In-situ electrochemical conversion of vanadium dioxide for enhanced zinc-ion storage with large voltage range. J. Power Sources 2021, 487, 229369. [Google Scholar] [CrossRef]
  76. Ding, J.; Du, Z.; Li, B.; Wang, L.; Wang, S.; Gong, Y.; Yang, S. Unlocking the potential of disordered rocksalts for aqueous zinc-ion batteries. Adv. Mater. 2019, 31, 1904369. [Google Scholar] [CrossRef]
  77. Cao, Z.; Zhang, H.; Ge, Y.; Clemente, R.; Dong, P.; Wang, L.; Shen, J.; Ye, M.; Ajayan, P.M. An in situ electrochemical oxidation strategy for formation of nanogrid-shaped V3O7·H2O with enhanced zinc storage properties. J. Mater. Chem. A 2019, 7, 25262. [Google Scholar] [CrossRef]
  78. Georgiou, T.; Jalil, R.; Belle, B.D.; Britnell, L.; Gorbachev, R.V.; Morozov, S.V.; Kim, Y.-J.; Gholinia, A.; Haigh, S.J.; Makarovsky, O.; et al. Vertical field-effect transistor based on graphene–WS2 heterostructures for flexible and transparent electronics. Nat. Nanotechnol. 2013, 8, 100–103. [Google Scholar] [CrossRef] [PubMed]
  79. Braga, D.; Lezama, I.G.; Berger, H.; Morpurgo, A.F. Quantitative determination of the band gap of WS2 with ambipolar ionic liquid–gated transistors. Nano Lett. 2012, 12, 5218–5223. [Google Scholar] [CrossRef] [Green Version]
  80. Debela, T.T.; Lim, Y.R.; Seo, H.W.; Kwon, I.S.; Kwak, I.H.; Park, J.; Cho, W.I.; Kang, H.S. Two-Dimensional WS2@Nitrogen-Doped graphite for high-performance lithium ion batteries: Experiments and molecular dynamics simulations. ACS Appl. Mater. Interfaces 2018, 10, 37928–37936. [Google Scholar] [CrossRef] [PubMed]
  81. Ratha, S.; Rout, C.S. Supercapacitor electrodes based on layered tungsten disulfide reduced graphene oxide hybrids synthesized by a facile hydrothermal method. ACS Appl. Mater. Interfaces 2013, 5, 11427–11433. [Google Scholar] [CrossRef] [PubMed]
  82. Liu, Q.; Li, X.; Xiao, Z.; Zhou, Y.; Chen, H.; Khalil, A.; Xiang, T.; Xu, J.; Chu, W.; Wu, X.; et al. Stable metallic 1T-WS2 nanoribbons intercalated with ammonia ions: The correlation between structure and electrical/optical properties. Adv. Mater. 2015, 27, 4837–4844. [Google Scholar] [CrossRef] [PubMed]
  83. Lv, R.T.; Robinson, J.A.; Schaak, R.E.; Sun, D.; Sun, Y.F.; Mallouk, T.E.; Terrones, M. Transition metal dichalcogenides and beyond: Synthesis, properties, and applications of single and few-layer nanosheets. Acc. Chem. Res. 2015, 48, 56. [Google Scholar] [CrossRef]
  84. Ma, Y.D.; Dai, Y.; Guo, M.; Niu, C.W.; Zhu, Y.T.; Huang, B.B. Evidence of the existence of magnetism in pristine VX2 Monolayers (X = S, Se) and their strain-induced tunable magnetic properties. ACS Nano 2012, 6, 1695. [Google Scholar] [CrossRef] [PubMed]
  85. Xu, K.; Chen, P.; Li, X.; Wu, C.; Guo, Y.; Zhao, J.; Wu, X.; Xie, Y. Ultrathin Nanosheets of Vanadium Diselenide: A Metallic Two-Dimensional Material with Ferromagnetic Charge-Density-Wave Behavior. Angew. Chem. Int. Ed. 2013, 52, 10477. [Google Scholar] [CrossRef]
  86. Yang, J.R.F.; Lv, F.; Zhou, J.H.; Lin, C.F.; Wang, K.; Zhang, Y.L.; Yang, Y.; Wang, W.; Li, J.B.; Guo, S.J. Metallic graphene-like VSe2 ultrathin nanosheets: Superior potassium-ion storage and their working mechanism. Adv. Mater. 2018, 30, 1800036. [Google Scholar] [CrossRef]
  87. Zhang, R.X.; Pan, C.S.; Nuzzo, R.G.; Gewirth, A.A. CoS2 as a sulfur redox-active cathode material for high-capacity nonaqueous Zn batteries. J. Phys. Chem. C 2019, 123, 8740. [Google Scholar] [CrossRef]
  88. Li, W.; Wang, K.L.; Cheng, S.J.; Jiang, K. An ultrastable presodiated titanium disulfide anode for aqueous “Rocking-Chair” zinc ion battery. Adv. Energy Mater. 2019, 9, 1900993. [Google Scholar] [CrossRef]
Figure 1. (a) Disassembly diagram of coin-type ZIBs; (b) The trend of publications on TMDs as ZIBs cathode materials.
Figure 1. (a) Disassembly diagram of coin-type ZIBs; (b) The trend of publications on TMDs as ZIBs cathode materials.
Materials 15 02654 g001
Figure 3. Defect atoms in TMDs singer layer.
Figure 3. Defect atoms in TMDs singer layer.
Materials 15 02654 g003
Figure 4. (a) HRTEM image of the defect−rich MoS2 and its rate performance [38]; (b) HRTEM images of oxygen incorporated MoS2 (left showing oxygen incorporated MoS2, right showing pristine MoS2) and the rate performance of oxygen incorporated MoS2 [40]; (c) HRTEM image of MoS2·nH2O and the rate performance of Zn/MoS2·nH2O batteries [41]; (d) HRTEM image of the MoS2/PANI−150 hybrid and its rate performance [42]; (e) SEM image and rate performance of the MoS2@CNTs hybrid electrode [43]; (f) HRTEM image of MoS2−160, and cycling performances at current densities of 1.0 A g−1 of MoS2 at various temperatures [44].
Figure 4. (a) HRTEM image of the defect−rich MoS2 and its rate performance [38]; (b) HRTEM images of oxygen incorporated MoS2 (left showing oxygen incorporated MoS2, right showing pristine MoS2) and the rate performance of oxygen incorporated MoS2 [40]; (c) HRTEM image of MoS2·nH2O and the rate performance of Zn/MoS2·nH2O batteries [41]; (d) HRTEM image of the MoS2/PANI−150 hybrid and its rate performance [42]; (e) SEM image and rate performance of the MoS2@CNTs hybrid electrode [43]; (f) HRTEM image of MoS2−160, and cycling performances at current densities of 1.0 A g−1 of MoS2 at various temperatures [44].
Materials 15 02654 g004
Figure 5. Ion Intercalation in single layer TMDs structure.
Figure 5. Ion Intercalation in single layer TMDs structure.
Materials 15 02654 g005
Figure 6. Modification of TMDs by Hybrid Engineering.
Figure 6. Modification of TMDs by Hybrid Engineering.
Materials 15 02654 g006
Figure 7. (a) HR−TEM image and rate performance of layered VS2 [45]; (b) SEM image of VS2 grown on SS mesh and rate performance of VS2@SS electrode [46]; (c) HRTEM image and rate performance of the rGO−VS2 composites [47]; (d) SEM image and rate performance of the VS2@VOOH−18h [48]; (e) HRTEM image and rate performance of VS2·NH3 electrode [49]; (f) TEM image and rate capability of in−situ electrochemical oxidation formed VS2/VOx [50].
Figure 7. (a) HR−TEM image and rate performance of layered VS2 [45]; (b) SEM image of VS2 grown on SS mesh and rate performance of VS2@SS electrode [46]; (c) HRTEM image and rate performance of the rGO−VS2 composites [47]; (d) SEM image and rate performance of the VS2@VOOH−18h [48]; (e) HRTEM image and rate performance of VS2·NH3 electrode [49]; (f) TEM image and rate capability of in−situ electrochemical oxidation formed VS2/VOx [50].
Materials 15 02654 g007
Figure 8. Modification of TMDs by in-situ electrochemical oxidation.
Figure 8. Modification of TMDs by in-situ electrochemical oxidation.
Materials 15 02654 g008
Figure 9. (a) HRTEM image, Cycling performances at current density of 200 mA g−1 of WS−200 [51]; (b) HRTEM image and rate performance of VSe2 [52].
Figure 9. (a) HRTEM image, Cycling performances at current density of 200 mA g−1 of WS−200 [51]; (b) HRTEM image and rate performance of VSe2 [52].
Materials 15 02654 g009
Table 1. Fabrication methods, precursors and synthesis conditions of TMDs as ZIBs cathode.
Table 1. Fabrication methods, precursors and synthesis conditions of TMDs as ZIBs cathode.
ProductsMethodPrecursorsTemperatureDurationRef
MoS2−xHydrothermal(NH4)6Mo7O24·4H2O, TAA200 °C18 h[38]
E- MoS2HydrothermalNa2MoO4, CS(NH2)2, carbon cloth, glucose, HCl190 °C24 h[39]
MoS2-OHydrothermal(NH4)6Mo7O24·4H2O, thiourea180 °C24 h[40]
MoS2·nH2OHydrothermal(NH4)6Mo7O24·4H2O, thiourea170 °C24 h[41]
MoS2/PANISolvothermalNa2MoO4, thiourea, C17H33CO2Na, ethanol, OA, HCl180 °C24 h[42]
MoS2@CNTsHydrothermalNa2MoO4·2H2O, thiourea, CNTs, glucose200 °C24 h[43]
MoS2-160Hydrothermal(NH4)6Mo7O24·4H2O, thiourea160 °C24 h[44]
VS2HydrothermalNH4VO3, TAA, NH3·H2O180 °C20 h[45]
VS2@SSHydrothermalNH4VO3, TAA, NH3·H2O, stainless steel mesh180 °C10 h[46]
rGO-VS2SolvothermalVO(acac)2, cysteine, GO, NMP200 °C8 h[47]
VS2@VOOHHydrothermalV2O5, TAA, NH3·H2O180 °C18 h[48]
VS2·NH3SolvothermalVO(acac)2, TAA, NMP200 °C24 h[49]
VS2/VOxSolvothermalNa3VO4·12H2O, TAA, ethylene180 °C20 h[50]
1T-WS2SolvothermalWCl6, TAA, DMF200 °C24 h[51]
VSe2Chemical Liquid Phase SynthesisVO(acac)2, Se powder, OAm330 °C5 h[52]
Table 2. Summary of the electrochemical performance of TMDs as ZIBs cathode.
Table 2. Summary of the electrochemical performance of TMDs as ZIBs cathode.
Cathode MaterialElectrolyteVoltageCapacity [mAh•g−1]Cycle StabilityRef
MoS2-O3M Zn (CF3SO3)20.2–1.4V232 at 0.1 A g−168% after 2000 cycles at 1.0 A g−1[40]
E- MoS22M Zn (CF3SO3)20.3–1.5V202.6 at 0.1 A g−198.6% after 600 cycles at 1.0 A g−1[39]
MoS2/PANI3M Zn (CF3SO3)20.2–1.3V181.6 at 0.1 A g−186% after 1000 cycles at 1.0 A g−1[42]
MoS2@CNTs3M Zn (CF3SO3)20.3–1.2V180 at 0.1 A g−180.1% after 500 cycles at 1.0 A g−1[43]
MoS2-1603M Zn (CF3SO3)20.25–1.25V168 at 0.1 A g−198.1% after 400 cycles at 1.0 A g−1[44]
MoS2·nH2O3M Zn (CF3SO3)20.2–1.25V165 at 0.1 A g−188% after 800 cycles at 2.0 A g−1[41]
MoS2−x3M Zn (CF3SO3)20.25–1.25V138.6 at 0.1 A g−187.8% after 1000 cycles at 1.0 A g−1[38]
VS2·NH32M Zn (CF3SO3)20.2–1.7V390 at 0.1 A g−1110% after 2000 cycles at 3.0 A g−1[49]
VS2/VOx25M ZnCl20.1–1.8V260 at 0.1 A g−175% after 3000 cycles at 1.0 A g−1[50]
rGO-VS23M Zn (CF3SO3)20.4–1.7V238 at 0.1 A g−193% after 1000 cycles at 5.0 A g−1[47]
VS2@SS1M ZnSO40.4–1.0V187 at 0.1 A g−180% after 2000 cycles at 2.0 A g−1[46]
VS2@VOOH3M Zn (CF3SO3)20.4–1.0V165 at 0.1 A g−186% after 200 cycles at 0.5 A g−1[48]
VS21M ZnSO40.4–1.0V159.1 at 0.1 A g−198% after 200 cycles at 0.5 A g−1[45]
WS-2001M ZnSO40.1–1.5V206.25 at 0.1 A g−10% after 100 cycles at 0.2 A g−1[51]
VSe22M ZnSO40.1–1.6V131.8 at 0.1 A g−180.8% after 500 cycles at 0.1 A g−1[52]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, T.; Li, H.; Yuan, J.; Xia, Y.; Liu, Y.; Sun, A. Recent Advance and Modification Strategies of Transition Metal Dichalcogenides (TMDs) in Aqueous Zinc Ion Batteries. Materials 2022, 15, 2654. https://doi.org/10.3390/ma15072654

AMA Style

Li T, Li H, Yuan J, Xia Y, Liu Y, Sun A. Recent Advance and Modification Strategies of Transition Metal Dichalcogenides (TMDs) in Aqueous Zinc Ion Batteries. Materials. 2022; 15(7):2654. https://doi.org/10.3390/ma15072654

Chicago/Turabian Style

Li, Tao, Haixin Li, Jingchen Yuan, Yong Xia, Yuejun Liu, and Aokui Sun. 2022. "Recent Advance and Modification Strategies of Transition Metal Dichalcogenides (TMDs) in Aqueous Zinc Ion Batteries" Materials 15, no. 7: 2654. https://doi.org/10.3390/ma15072654

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