Next Article in Journal
Reverse Magnetization Behavior Investigation of Mn-Al-C-(α-Fe) Nanocomposite Alloys with Different α-Fe Content Using First-Order Reversal Curves Analysis
Next Article in Special Issue
Recent Advances in g-C3N4 for the Application of Perovskite Solar Cells
Previous Article in Journal
Fitting Procedure to Reconstruct the Size Distribution and the Concentration of Silver Colloidal Nanoparticles from UV-Vis Spectra
Previous Article in Special Issue
Graphene-Modified Co-B-P Catalysts for Hydrogen Generation from Sodium Borohydride Hydrolysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances of Transition Metal Chalcogenides as Cathode Materials for Aqueous Zinc-Ion Batteries

School of Materials Science and Engineering, Shenyang University of Technology, Shenyang 110870, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(19), 3298; https://doi.org/10.3390/nano12193298
Submission received: 4 September 2022 / Revised: 17 September 2022 / Accepted: 20 September 2022 / Published: 22 September 2022

Abstract

:
In recent years, advances in lithium-ion batteries (LIBs) have pushed the research of other metal-ion batteries to the forefront. Aqueous zinc ion batteries (AZIBs) have attracted much attention owing to their low cost, high capacity and non-toxic characteristics. Among various cathodes, transition metal chalcogenides (TMCs) with a layered structure are considered as suitable electrode materials. The large layer spacing facilitates the intercalation/de-intercalation of Zn2+ between the layers. In this mini-review, we summarize a variety of design strategies for the modification of TMCs. Then, we specifically emphasize the zinc storage capacity of the optimized electrodes. Finally, we propose the challenges and future prospects of cathode materials for high-energy AZIBs.

1. Introduction

The growing energy crisis has driven the unprecedented development of renewable clean energy [1,2,3]. To date, lithium-ion batteries (LIBs) are the most widely used energy storage devices. However, the scarcity of lithium resources, the inflammability of electrolytes, and high operating environment requirements limit their growth [4,5,6,7]. Rechargeable aqueous zinc ion batteries (AZIBs) are a new generation of safety batteries. They possess certain advantages in terms of abundant zinc reserves, low anode potential (−0.763 V vs. SHE) and high theoretical capacity (820 mAh g−1) [8,9,10,11]. Therefore, AZIBs have become one of the candidates to replace LIBs. However, zinc anodes undergo dissolution–precipitation reactions with several adverse reactions, such as dendrite growth, corrosion, and by-product formation. They are inevitable during repeated plating and stripping and seriously damage the cycle life of the cells [12]. Also, divalent zinc ions possess stronger electrostatic interactions than monovalent lithium ions [13,14]. Therefore, the choice of a suitable intercalation material is the crucial to break through this challenge.
In recent years, many efforts are devoted to the exploration of cathode materials, including Prussian blue analogs, vanadium-based and manganese-based compounds, and transition metal chalcogenides (TMCs) [15,16,17,18,19]. Among them, Prussian blue analogs are featured by a high voltage window, but their crystal structure is unstable and prone to phase transformation [20]. The inherent low electrical conductivity and poor structural stability of V- and Mn-based materials lead to their slow electrochemical kinetics [21,22,23]. The electrical conductivity of TMCs is superior to that of oxides. Additionally, TMCs are characterized by a unique layer structure with large layer spacing. Their high specific surface area can provide many active sites and reduce ion transfer paths [24]. In previous reports, Naveed et al. designed the VS2 nanosheet materials as cathodes, which maintain a capacity of 138.3 mAh g−1 at 0.1 A g−1 after 500 cycles with a retention of 94.38% [25]. Kang’s group summarized the Zn storage ability of various oxides, sulfides and borides [26]. The results show that the activated MnS electrode is a potential cathode material with both high capacity and stable cycling performance. However, bulk MoS2 and WS2 materials are virtually incapable of storing zinc ions. Therefore, it is essential to improve the electrochemical performance of TMCs by effective tuning strategies.
In recent years, there have been numerous reports on TMCs-based cathode materials. Herein, we first summarize several feasible strategies for optimizing the electrode structure. Then, we discuss the electrochemical performance of TMCs cathodes for AZIBs. Lastly, we overview the advances in cathode materials and present the current challenges and prospects for constructing advanced cathodes of AZIBs.

2. The Electrochemical Performance of TMCs Cathodes

Two-dimensional (2D) TMCs are composed of transition metals (M = V, Ni, Mo, W and Mn, etc.) and chalcogen elements (X = S, Se, Te) with tunable electrical properties from semiconductors to metals. They is widely studied in the field of energy storage and conversion field [24,27,28]. In addition, graphene-like 2D layered TMCs are highly advantageous for battery applications because of their large specific surface area, which can significantly increase the contact area between the active material and the electrolyte [29]. Figure 1 illustrates the crystal structures of various TMCs materials. Their non-bonding properties allow the insertion of atoms, ions and molecules. The design strategies of TMCs materials mainly include defect engineering, hybridization, phase modulation, and in situ electrochemical oxidation. The main focus is to widen the interlayer space, improve the electrode conductivity, and accelerate the electrochemical kinetic process. We will categorize the strategies of structural design and zinc storage capabilities of various cathode materials in the following sections.

2.1. VS2 and VS4

The hexagonal-structured vanadium disulfide (VS2) is a member of the TMC family. It owns a layered structure with a layer spacing of 5.76 Å [30]. The S-V-S layers rely on weak van der Waals interactions (Figure 2a). Hence, VS2 materials have become an attractive host for Zn ion insertion/extraction. Nevertheless, VS2 is unstable in aqueous solutions, leading to severe capacity decay during cycling. From the aspect of modulating the material structure, coating is considered to be effective strategy. Pu et al. prepared rose-shaped VS2 encapsulated with a hydrophilic VOOH coating using a one-pot hydrothermal route [31]. The assembled cells maintained 82% capacity after 400 cycles. The rate capability and long cycle life of the optimized sample is significantly improved compared to the VS2 sample, which is attributed to the O–H in VOOH. From Figure 2b, its presence not only enhances the wetting of the electrode and electrolyte, but also prevents the dissolution of the main materials. Fan’s group prepared ultrathin VS2 nanosheets grown on graphene sheets (rGO–VS2) by a solvothermal strategy [32]. The rGO offers a large specific surface area and excellent electrical conductivity. In addition, the close contact of VS2 nanosheets with rGO can effectively prevent the dissolution and corrosion of the host materials. It ensures the high stability of the electrode in long-term cycling. Thus, Zn/rGO–VS2 cells can deliver a large specific capacity (238 mAh g−1 at 0.1 A g−1) and excellent rate performance (190 mAh g−1 at 5 A g−1, Figure 2c). After 1000 cycles of charging and discharging, it can still maintain 93.3% of the initial capacity, as shown in Figure 2d.
The N dopant provides high affinity for the transition metals, so it is possible to form a strong coupling between the host material and the N-doped carbon. This contributes to accelerate the interfacial electron transfer and reduces cycling-induced stress and volume changes. Liu and co-workers prepared spun VS2 materials on a N-doped carbon layer (VS2@N-C) by an in situ hybridization strategy [33]. This strategy ensures a strong interfacial interaction between the active material and the N-doped carbon. It promotes the enhancement of electrochemical kinetics and cycling stability of the cathode. The optimal electrode possesses a capacity of 203 mAh g−1 at 0.05 A g−1. Based on the Zn ion insertion and extraction mechanism, the cells can obtain a capacity of 144 mAh g−1 after 600 cycles. Liu’s group synthesized 1T-VS2 colloidal nanospheres assembled from nanoflakes [34]. By controlling the charge cutoff voltage, a number of the Zn ions were trapped in the interlayer of the structure after the initial charge/discharge cycle. These “dead Zn” act as “pillars” to ensure the stability of the layered structure of VS2. After 2000 cycles, the cells maintained the capacity retention of 86.7%. This unique layered structure increases the conductivity due to the presence of carbon and oxygen groups on the surface, facilitating the penetration of the aqueous electrolyte and providing more active sites.
In order to further improve the stability of VS2 electrodes in an aqueous electrolyte. Yang et al. employed an in situ electrochemical oxidation approach to enhance the interlayer space of vanadium disulfide (VS2NH3) hollow spheres [35]. This large layer spacing (1.21 nm) is favorable to the improvement of zinc ion storage capacity. The VS2NH3 samples transform into a porous structured V2O5·nH2O phase during the first charging cycle (Figure 2e). It enhances the active sites and is conducive to a rapid electrochemical kinetic process. This derived electrode maintains a high capacity at 3 A g−1 even after 2000 cycles. Similarly, Du and co-workers proposed the formation of VS2/VOx heterostructures by in situ electrochemical induction [36]. When the sample is charged to 1.8 V, the morphology of the composite changes from rose-like shape to sheet-like one. This structure can withstand the volume expansion caused by repeated cycles. Compared to the pure Zn–VS2 cell, the Zn–VS2/VOx one demonstrates a cycling stability of 3000 cycles at 1 A g−1 with an improved working potential of 0.25 V (Figure 3a). This strategy of combining highly conductive sulfides and excellent chemically stable oxides leads to an enhancement of the Zn2+ storage capacity of the VS2/VOx cathode.
VS4 is a material with a one-dimensional (1D) atomic chain structure. When compared with the VS2 material, it possesses many S atoms with layer spacing up to 5.83 Å, as shown in Figure 3b. It indicates that VS4 may show excellent zinc storage capacity. Zhu et al. prepared VS4 materials as cathodes via a hydrothermal route [37]. Density functional theory (DFT) calculations demonstrate that the electrode is capable of storing zinc ions up to a maximum specific capacity of 262 mAh g−1. Then, the cell can deliver a specific capacity of 310 mAh g−1 at 0.1 A g−1. It is higher than the theoretical capacity, which could be due to the additional absorption capacity. Furthermore, the absence of additional by-product generation suggests that the energy storage of material follows a zinc-ion embedding/de-embedding mechanism. The construction of heterostructures is also an attractive strategy. In theory, the intrinsic zinc storage capacity of the cathode can be effectively enhanced by building a heterostructure with sufficient interfaces and grain boundaries. Fang’s group designedVS4/V2O3 heterostructures with a high specific surface area [38]. The assembled battery shows a capacity of 163 mAh g−1 at 0.1 A g−1. As a contrast, the single electrode presents inferior electrochemical performance. This demonstrates that the optimized heterogeneous material can boost the energy storage capacity.
It is also an effective strategy for improving electrochemical kinetics by compositing with highly conductive materials. Qin and co-workers synthesized VS4 composite material immobilized on reduced graphene oxide (VS4@rGO) as a cathode [39]. This synergistic effect enables the VS4@rGO electrode to reach a capacity of 180 mAh g−1 (1 A g−1) after 165 cycles with a capacity retention of 93.3%. Nevertheless, the above-mentioned electrochemical performance is still unsatisfactory. Chen and co-workers optimized the morphology of the VS4@rGO composites to achieve a specific capacity of 450 mAh g−1 at a current density of 0.5 A g−1 when used as cathodes [40]. Moreover, the capacity of 313.8 mAh g−1 was maintained at high current densities (10 A g−1). It indicates that the batteries possess an excellent rate capability. It is noteworthy that a new phase Zn3(OH)2V2O7·2H2O (ZVO) appears during charging. The following reactions may occur in the electrode material:
VS4 + xZn2+ + 2xe ↔ ZnxVS4
VS4 + 11H2O + 3Zn2+ → Zn3(OH)2V2O7·2H2O + 8S + 16H+ + 10e
After that, it transforms into the ZnV3O8 phase during the long cycles, which is associated with the subsequent capacity decay. Gao et al. reported a flower-like VS4/CNTs cathode with an abundant mesoporous structure, which effectively shortens the diffusion path of zinc ions [41]. In Figure 3c, when the first cycle is charged to 1.7 V, the charging curve undergoes a slow upward trend, which implies a phase transition process. Figure 3d further confirms that the mechanism of the phase change reaction of VS4 with zinc pyrovandate (Zn3+x(OH)2V2O7·2H2O). The results show that the Zn–VS4/CNTs batteries possess a reversible capacity of 265 mAh g−1 (0.25 A g−1) and a good rate performance in the potential range from 0.2 to 1.7 V. Although the energy storage capacity has been significantly improved by modification of the electrode material, the inevitable phase change during the reaction process still hinders the cycle life. This may be related to the high charging voltage.

2.2. MoS2

MoS2 is a typical 2D-layered structure bound by weak van der Waals forces [42,43,44]. However, the ionic radius of hydrated Zn2+ is 0.43 nm, which places high demands on the interlayer space of the host materials. To enhance the reaction kinetics of Zn ion insertion and extraction, Li et al. extended the interlayer spacing of the (002) plane of MoS2 nanosheets from 0.62 nm to 0.70 nm [45]. Due to the addition of glucose, an amorphous carbon layer is wrapped on the surface of MoS2. This facilitates the alleviation of volume expansion and promotes charge transfer. From Figure 4a, the specific capacity of the batteries can be maintained at 164.5 mAh g−1 after 600 cycles. In Figure 4b, the charge storage mechanism can be described as follow:
Cathode: xZn2+ + x2e + MoS2 ↔ ZnxMoS2
Anode: Zn2+ + 2e ↔ Zn
In addition, a flexible solid-state Zn/E-MoS2 cell was further assembled using the starch/polyacrylamide (PAM) polymer electrolyte. Under different mechanical strengths, the cell still can maintain a stable charge/discharge process.
Due to the diversity of coordination of Mo and S atoms, MoS2 can show a semiconductor phase with a triangular prismatic structure (2H phase) and a metallic phase with an octahedral structure (1T phase). 1T-phase MoS2 possesses higher electrical conductivity and better hydrophilicity than 2H-phase ones [46,47]. Therefore, the material is also a promising electrode for zinc storage. Huang et al. synthesized a 1T-phase MoS2 nanosheet grown directly on reduced graphene oxide (rGO) scaffolds [48]. The addition of the rGO scaffold can serve to stabilize the 1T phase and reduce the possibility of phase transition during zinc ions insertion/extraction. In addition, it can improve the electrical conductivity, thus shortening the diffusion path of zinc ions. The initial discharge capacity of the 1T-MoS2/rGO heterogeneous electrode is 108.3 mAh g−1, and the cell maintains a capacity retention of 88% after repeated charge/discharges of 1000 times.
Tang’s group synthesized N-doped 1T MoS2 nanoflowers assembled from ultrathin nanosheets by a one-step hydrothermal sulfidation of Mo-based organic framework (MOF) precursors [49]. The introduction of defects effectively widens the interlayer spacing and increases the number of sulfur vacancies as well as the hydrophilicity of the sample. Zn/N-doped 1T MoS2 batteries deliver a capacity of 149.6 mAh g−1 (0.1 A g−1). The capacity retention is up to 89.1% after 1000 cycles at 3 A g−1. Additionally, the electrochemical performance was studied for the difference in area mass loading of the electrodes. The area capacity shows an outstanding performance when the area loading reaches 1.701 mg cm−2. This implies that the electrode presents excellent rate capacity even at high loadings. Liu and co-workers reported MoS2 nanosheets with different phase contents as cathode materials [50]. Among them, the MoS2 nanosheet electrode with 1T phase of content around 70% presents favorable long-term cycling stability. This indicates that the presence of metallic 1T phase favors the ion and charge transfer.
Apart from composite with conductive materials, combination with organic molecules also promotes the increase of zinc ion storage capacity. Yao et al. designed a 2D MoS2/C19H42N+ (CTAB) organic–inorganic superlattice structure (MoS2–CTAB) as a cathode [51]. This unique structure can significantly enlarge the interlayer spacing (1.0 nm) of the host materials (Figure 4c). In addition, the stable electrode structure can accommodate the expansion and contraction of Zn2+ within the host structure. The loading mass of the active material is a very important parameter for the evaluation of the specific capacity and energy density of the cell. The areal capacity of the battery increases with the area mass loading in a potential window of 0.2–1.3 V. Figure 4d demonstrates the optimal adsorption position of Zn ions at the pure MoS2 and modified MoS2 electrodes by DFT calculations. In the former structure, the Zn ion prefers to adsorb at the top site of the Mo atom with a corresponding energy of −0.31 eV, but the charge accumulation of adsorbed Zn with adjacent S atoms suggests a strong electrostatic interaction between Zn and the host material. In the latter one, the adsorption energy of Zn ion at the same site is −0.25 eV, and the distance between Zn and its three neighboring S atoms remains almost unchanged relative to the original electrode.

2.3. MnS

In recent years, many efforts have been made in Mn-based oxide cathodes [52]. For instance, Minakshi et al. compared the cathodic behavior of electrolytic manganese dioxide (EMD) and chemically prepared battery-grade manganese dioxide (BGM) in a lithium hydroxide (LiOH) electrolyte [53]. The EMD cell demonstrated stable discharge/charge cycles compared to the BGM. Wang’s group designed nanocrystal line structures of MnO2 materials with particle sizes typically less than 10 nm [54]. This structure confers some electrode/electrolyte contact interfaces. Therefore, the Zn/MnO2 cell delivers a capacity of 260 mAh g−1 at 1.3 C.
However, their rate capability and cycle stability cannot meet the current high-capacity energy storage requirements. In addition, Zn ions show strong electrostatic interactions with the Mn oxide lattice, leading to large energy barriers for Zn2+ migration [55]. Chen et al. reported the transformation of α-MnS materials into high-performance manganese oxide cathodes (MnS–EDO) by in situ electrochemical oxidation [56]. Compared to α-MnO2, this electrode generated more defects and vacancies after structural reconfiguration. This indicates a rapid electrochemical kinetic process. The Zn/MnS–EDO cell shows a high specific capacity of 335.7 mAh g−1 with capacity retention close to 100% and reversible rate performance (Figure 5a). In addition, it can undergo a repeated charge/discharge process of 4000 cycles, as shown in Figure 5b.
To further enhance the electrical conductivity of MnS, Ma and co-workers designed a MnS and rGO composite material. This synergistic effect effectively improves the Zn storage capacity of the electrode material [57]. MnS possesses various phase types: α-MnS, β-MnS and γ-MnS. Both β- and γ-phases are sub-stable and readily transform to the stable rock salt structure α-MnS (Figure 5c). Jiang et al. fabricated flexible zinc ion microcells with MnS as cathodes and guar gels as the quasi-solid electrolyte by etching soft templates on various substrates [58]. The cells prepared on PET substrates deliver an area-specific capacity of 178 μAh cm−2. After 1000 cycles, they can maintain a capacity of 150 mAh g−1 at 1 A g−1 in Figure 5d. Moreover, the area energy density can reach 322 μWh cm−2 at a power density of 120 μW cm−2. In addition, this quasi-solid-state cell shows excellent flexibility with almost no significant capacity degradation when the device is bent at multiple angles.

2.4. VSe2

Among numerous TMCs, vanadium diselenide (VSe2) presents a typical layered structure with a sandwich-like Se–V–Se connected by van der Waals interactions [59]. The materials possess a layer spacing of 6.11 Å, which can provide sufficient transport channels and active sites for the intercalated ions [60]. Moreover, the strong electronic coupling force between adjacent V4+–V4+ endows it with metallic properties. Thus, it shows great potential in terms of energy storage [61,62]. For instance, Alshareef’s group explored the energy storage capacity of VSe2 materials in different alkali metal batteries [63]. The morphology of VSe2 nanosheets can be modulated using N-methylpyrrolidone (NMP) solvent, and the electrochemical performance of the samples is further improved by in situ carbon coating. The electrodes can provide a specific capacity of 768 mAh g−1 (lithium storage) and 571 mAh g−1 (sodium storage), respectively.
Recently, VSe2 materials also show attractive performance in zinc storage. Wu et al. designed an ultrathin VSe2 nanosheet as a cathode [64]. The assembled cells maintain an initial capacity of 80.8% after 500 cycles. This durable cycling stability is attributed to their fast Zn2+ diffusion kinetic process and durable cycling stability. The local charge density map in Figure 6a shows a decrease in charge around the Zn ion and an increase in charge around the Se and some V sites. This demonstrates that the intercalated Zn is bonded to the Se ligand. From Figure 6b, the migration barrier for the optimal diffusion pathway of Zn ions is 0.91 eV, which corresponds to a fast ion migration rate. Based on the Zn ion insertion and extraction mechanism (Figure 6c), the cell possesses an energy density of 107.3 Wh kg−1 at a power density of 81.2 W kg−1. Cai and co-workers synthesized homogeneous flower-shaped VSe2 spheres using MXene as a support [65]. The specific capacity of Zn–VSe2/MXene cells was higher than the initial capacity after 2000 cycles at 5 A g−1 in the voltage window of 0.2–1.6 V (Figure 6d). This may be attributed to the generation of a Zn0.25V2O5 (ZVO) phase. With repeated discharge/charging, the generation and accumulation of ZVO phase provides a continuous capacity contribution to the battery.

2.5. Ni3S2

Nickel sulfide (Ni3S2) is a promising electrode material with the advantages of high activity and theoretical capacity, excellent stability, and low cost [66,67]. However, Ni3S2 materials perform inferiorly in cycling stability and specific capacity when they are used as a cathode for AZIBs. Structural defects are an approach to modulate the crystal structure of the materials [68]. In addition, the introduction of defects can increase the electrical conductivity and carrier density, effectively improving the electrochemical activity of the electrode materials [69].
For instance, Tong et al. doped highly reactive Co ions in Ni3S2 nanocones by atomic layer deposition (ALD) and the hydrothermal method [70]. A Co12-Ni3S2/NF (C12NS) sample was used as the cathode; the hydrogel electrolyte is immersed in 5 M KOH/0.1 M Zn(AC)2 solution to assemble a quasi-solid-state flexible cell. The reactivity of the electrodes is increased due to the Co doping and sulfation. The specific capacity of the activated electrode is about four times higher than that before activation. This is mainly due to the adsorption of OH radicals from the alkaline electrolyte on the sulfur sites of the Co-doped Ni3S2. The reaction mechanism can be described as:
Cathode: Co-Ni3S2 + 2OH ↔ Co-Ni3S2(OH)2 + 2e
Co-Ni3S2(OH)2 + 3OH ↔ Co-Ni3S2O2OH + 2H2O + 3e
Anode: Zn[(OH)4]2– + 2e ↔ Zn + 4OH
During the continuous adsorption of hydroxide radicals, sulfur and oxygen coexist. The formed oxygen–sulfur bonds enhance the electrochemical performance of the electrode material, as shown in Figure 7a. The Co-doped electrodes present a low energy barrier. It indicates that they can achieve a fast Zn ion migration (Figure 7b). In Figure 7c, the capacity retention of Co12-Ni3S2/NF electrode after 5000 cycles is up to 90%.
In addition to the defect strategy, Wang’s group prepared Ni/Ni3S2 nanocomposites with large specific surface areas using Ni–ZIF MOFs as precursors through a simple medium temperature solid–gas phase reaction [71]. The assembled batteries with 6 M KOH-0.6 M ZnO electrolyte possess an excellent rate capability (Figure 7d). They can maintain a capacity retention of 83% after 1000 cycles at 20 A g−1 (Figure 7e). Additionally, an energy density of 379 Wh kg−1 can be obtained at 340 W kg−1 at a high output voltage (1.7 V). Table 1 summarizes the electrochemical performance of TMC cathodes. It can be observed that these materials present some disadvantages, mainly in terms of limited voltage window, low specific capacity and inferior cycle life. In addition, the modification tactics lead to a large difference in the electrochemical performance of the electrodes. This demonstrates that an appropriate strategy can enhance the structural stability as well as the cycle life of the cells. The VS4 composite shows significant advantages with a cycle life of 3500 cycles with a stable charge/discharge process. It can be noted that the electrolytes used in these batteries are varied. The high concentration of electrolyte contributes to the increase in capacity. Therefore, the choice of electrolyte is also crucial in future work to affect the electrochemical performance of the Zn/TMCs batteries.

3. Summary and Outlook

AZIBs have attracted a considerable attention as an alternative to LIBs with the features of being inexpensive, environmentally friendly, resource-rich, and having high theoretical capacity. However, current cathode materials are still hampered by their inferior conductivity, slow kinetics, structural instability, and dissolution of active substances. Among them, TMCs possess a unique layered structure and a large layer spacing and high conductivity, which facilitate the transfer of ion carriers. First, the high specific surface area provides many electrochemically active sites and short ion transfer paths. Secondly, its excellent electrical conductivity ensures fast electron transfer. Finally, the open-layer structure favors the embedding of electrolyte ions and reduces the volume change. We summarized the research advances of TMCs as electrode materials in recent years. The main focus was on their modification strategies and the improvement of zinc storage capacity. Reasonable modification strategies are beneficial for the improvement of energy storage capability. Nevertheless, the goal of commercialization is still not reached. More efforts may be required to try to shorten this gap.
The electrochemical performance of TMCs materials, such as specific capacity and long cycle performance, still needs to be improved. There are several strategies for considering directions for this: (1) introducing defects (vacancies or doping) in electrode materials. It can add many active sites and conductivity to effectively increase the capacity of the battery. (2) Combining high conductivity and specific surface area materials such as MOFs and rGO. MOFs possess rich pore structures. The calcination of MOFs precursors is an effective method to prepare metal compound/carbon composites. (3) Expanded layer spacing. The increase of layer spacing is beneficial to the rapid shuttling of zinc ions, which can effectively improve the electrochemical kinetics.
Although some TMCs have been employed for energy storage, there are few materials with desired zinc storage capability. Therefore, it is still necessary to explore the unexploited cathode materials (Figure 8). We should place our eyes on the study of materials such as WS2, MoSe2, etc., and focus on the design of effective modification strategies. Additionally, the mechanism of electrochemical energy storage is still unclear. Advanced in situ characterization techniques are beneficial to explore the evolution of the structure of electrode materials during charging and discharging as well as elemental valence changes.
In addition, the structural stability of electrode materials is closely associated with their electrochemical performance. The cycling process may generate irreversible phase changes and side reactions, which lead to structural collapse or capacity loss. The modulation of heterogeneous interfaces may suppress the occurrence of side reactions. Finally, flexible electronic devices have attracted much attention because of their high safety, portability, and wearable features. There are few reports related to flexible devices for AZIBs. They should focus on flexible substrates and cathodes with certain mechanical flexibility, and stable solid-state electrolytes with high ionic conductivity.

Author Contributions

Conceptualization, Y.L.; methodology, Y.L.; software, Y.L.; validation, Y.L.; formal analysis, Y.L.; investigation, Y.L.; resources, X.W.; data curation, Y.L. and X.W.; writing—original draft preparation, Y.L.; writing—review and editing, X.W.; visualization, Y.L.; supervision, X.W.; project administration, X.W.; funding acquisition, X.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (No. 52172218).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Goodenough, J.B. Energy storage materials: A perspective. Energy Storage Mater. 2015, 1, 158–161. [Google Scholar] [CrossRef]
  2. Zhao, D.P.; Dai, M.Z.; Zhao, Y.; Liu, H.Q.; Liu, Y.; Wu, X. Improving electrocatalytic activities of FeCo2O4@FeCo2S4@PPy electrodes by surface/interface regulation. Nano Energy 2020, 72, 104715. [Google Scholar] [CrossRef]
  3. Ruan, P.; Liang, S.; Lu, B.; Fan, H.J.; Zhou, J. Design strategies for high-energy-density aqueous zinc batteries. Angew. Chem. Int. Ed. 2022, 61, e202200598. [Google Scholar] [CrossRef]
  4. Liu, C.; Wu, X.; Wang, B. Performance modulation of energy storage devices: A case of Ni-Co-S electrode materials. Chem. Eng. J. 2020, 392, 123651. [Google Scholar] [CrossRef]
  5. Tarascon, J.M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. [Google Scholar] [CrossRef]
  6. Yang, D.; Zhou, Y.; Geng, H.; Liu, C.; Lu, B.; Rui, X.; Yan, Q. Pathways towards high energy aqueous rechargeable batteries. Coord. Chem. Rev. 2020, 424, 213521. [Google Scholar] [CrossRef]
  7. Ju, Z.; Zhao, Q.; Chao, D.; Hou, Y.; Pan, H.; Sun, W.; Yuan, Z.; Li, H.; Ma, T.; Su, D.; et al. Energetic Aqueous Batteries. Adv. Energy Mater. 2022, 12, 2201074. [Google Scholar] [CrossRef]
  8. Liu, Y.; Liu, Y.; Wu, X. Toward long-life aqueous zinc ion batteries by constructing stable zinc anodes. Chem. Rec. 2022, e202200088. [Google Scholar] [CrossRef]
  9. Li, C.; Xie, X.; Liang, S.; Zhou, J. Issues and future perspective on zinc metal anode for rechargeable aqueous zinc-ion batteries. Energy Environ. Mater. 2020, 3, 146–159. [Google Scholar] [CrossRef]
  10. Liu, Y.; Wu, X. Hydrogen and sodium ions co-intercalated vanadium dioxide electrode materials with enhanced zinc ion storage capacity. Nano Energy 2021, 86, 106124. [Google Scholar] [CrossRef]
  11. Song, M.; Tan, H.; Chao, D.; Fan, H.J. Recent advances in Zn-ion batteries. Adv. Funct. Mater. 2018, 28, 1802564. [Google Scholar] [CrossRef]
  12. Yang, J.; Yin, B.; Sun, Y.; Pan, H.; Sun, W.; Jia, B.; Zhang, S.; Ma, T. Zinc Anode for Mild Aqueous Zinc-Ion Batteries: Challenges, Strategies, and Perspectives. Nano-Micro Lett. 2022, 14, 42. [Google Scholar] [CrossRef]
  13. Liu, Y.; Liu, Y.; Wu, X.; Cho, Y.R. Enhanced electrochemical performance of Zn/VOx batteries by a carbon-encapsulation strategy. ACS Appl. Mater. Interfaces 2022, 14, 11654–11662. [Google Scholar] [CrossRef]
  14. Li, H.F.; Ma, L.T.; Han, C.P.; Wang, Z.F.; Liu, Z.X.; Tang, Z.J.; Zhi, C.Y. Advanced rechargeable zinc-based batteries: Recent progress and future perspectives. Nano Energy 2019, 62, 550–587. [Google Scholar] [CrossRef]
  15. Liu, Y.; Wu, X. Review of vanadium-based electrode materials for rechargeable aqueous zinc ion batteries. J. Energy Chem. 2021, 56, 223–237. [Google Scholar] [CrossRef]
  16. Guo, X.; Li, J.; Jin, X.; Han, Y.; Lin, Y.; Lei, Z.; Wang, S.; Qin, L.; Jiao, S.; Cao, R. A hollow-structured manganese oxide cathode for stable Zn-MnO2 batteries. Nanomaterials 2018, 8, 301. [Google Scholar] [CrossRef]
  17. Liu, B. Transition metal dichalcogenides for high-performance aqueous zinc ion batteries. Batteries 2022, 8, 62. [Google Scholar] [CrossRef]
  18. Trócoli, R.; La Mantia, F. An aqueous zinc-ion battery based on copper hexacyanoferrate. ChemSusChem 2015, 8, 481–485. [Google Scholar] [CrossRef]
  19. Liu, Y.; Hu, P.F.; Liu, H.Q.; Wu, X.; Zhi, C.Y. Tetragonal VO2 hollow nanospheres as robust cathode material for aqueous zinc ion batteries. Mater. Today Energy 2020, 17, 100431. [Google Scholar] [CrossRef]
  20. Yang, Q.; Mo, F.; Liu, Z.; Ma, L.; Li, X.; Fang, D.; Chen, S.; Zhang, S.; Zhi, C. Activating C-coordinated iron of iron hexacyanoferrate for Zn hybrid-ion batteries with 10000-cycle lifespan and superior rate capability. Adv. Mater. 2019, 31, 1901521. [Google Scholar] [CrossRef]
  21. Liu, Y.; Liu, Y.; Yamauchi, Y.; Alothman, Z.A.; Kaneti, Y.V.; Wu, X. Enhanced zinc ion storage capability of V2O5 electrode materials with hollow interior cavities. Batt. Supercap. 2021, 4, 1867–1873. [Google Scholar] [CrossRef]
  22. Zhong, Y.; Xu, X.; Veder, J.P.; Shao, Z. Self-recovery chemistry and cobalt-catalyzed electrochemical deposition of cathode for boosting performance of aqueous zinc-ion batteries. iScience 2020, 23, 100943. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, Y.; Liu, Y.; Wu, X.; Cho, Y.R. High performance aqueous zinc battery enabled by potassium ion stabilization. J. Colloid Interface Sci. 2022, 628, 33–40. [Google Scholar] [CrossRef] [PubMed]
  24. Tan, C.; Cao, X.; Wu, X.J.; He, Q.; Yang, J.; Zhang, X.; Chen, J.; Zhao, W.; Han, S.; Nam, G.H.; et al. Recent advances in ultrathin two-dimensional nanomaterials. Chem. Rev. 2017, 117, 6225–6331. [Google Scholar] [CrossRef]
  25. Naveed, A.; Yang, H.; Shao, Y.; Yang, J.; Yanna, N.; Liu, J.; Shi, S.; Zhang, L.; Ye, A.; He, B.; et al. A Highly reversible Zn anode with intrinsically safe organic electrolyte for long-cycle-life batteries. Adv. Mater. 2019, 31, 1900668. [Google Scholar] [CrossRef]
  26. Liu, W.; Hao, J.; Xu, C.; Mou, J.; Dong, L.; Jiang, F.; Kang, Z.; Wu, J.; 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]
  27. Liu, J.; Zhou, W.; Zhao, R.; Yang, Z.; Li, W.; Chao, D.; Qiao, S.Z.; Zhao, D. Sulfur-based aqueous batteries: Electrochemistry and strategies. J. Am. Chem. Soc. 2021, 143, 15475–15489. [Google Scholar] [CrossRef]
  28. Gao, M.R.; Xu, Y.F.; Jiang, J.; Yu, S.H. Nanostructured metal chalcogenides: Synthesis, modification, and applications in energy conversion and storage devices. Chem. Soc. Rev. 2013, 42, 2986–3017. [Google Scholar] [CrossRef]
  29. Gong, J.; Li, H.; Zhang, K.; Zhang, Z.; Cao, J.; Shao, Z.; Tang, C.; Fu, S.; Wang, Q.; Wu, X. Zinc-ion storage mechanism of polyaniline for rechargeable aqueous zinc-ion batteries. Nanomaterials 2022, 12, 1438. [Google Scholar] [CrossRef]
  30. 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]
  31. 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 Source 2019, 437, 226917. [Google Scholar] [CrossRef]
  32. Chen, T.; Zhu, X.; 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 Source 2020, 477, 228652. [Google Scholar] [CrossRef]
  33. 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]
  34. Tan, Y.; Li, S.; Zhao, X.; Wang, Y.; Shen, Q.; Qu, X.; Liu, Y.; Jiao, L. Unexpected role of the interlayer “dead Zn2+” in strengthening the nanostructures of VS2 cathodes for high-performance aqueous Zn-ion storage. Adv. Energy Mater. 2022, 12, 2104001. [Google Scholar] [CrossRef]
  35. 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]
  36. 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. 2020, 31, 2008743. [Google Scholar] [CrossRef]
  37. Zhu, Q.; Xiao, Q.; Zhang, B.; Yan, Z.; Liu, X.; Chen, S.; Ren, Z.; Yu, Y. VS4 with a chain crystal structure used as an intercalation cathode for aqueous Zn-ion batteries. J. Mater. Chem. A 2020, 8, 10761–10766. [Google Scholar] [CrossRef]
  38. Ding, J.; Gao, H.; Liu, W.; Wang, S.; Wu, S.; Fang, S.; Cheng, F. Operando constructing vanadium tetrasulfide-based heterostructures enabled by extrinsic adsorbed oxygen for enhanced zinc ion storage. J. Mater. Chem. A 2021, 9, 11433–11441. [Google Scholar] [CrossRef]
  39. Qin, H.; Yang, Z.; Chen, L.; Chen, X.; Wang, L. A high-rate aqueous rechargeable zinc ion battery based on the VS4@rGO nanocomposite. J. Mater. Chem. A 2018, 6, 23757–23765. [Google Scholar] [CrossRef]
  40. Chen, K.; Li, X.; Zang, J.; Zhang, Z.; Wang, Y.; Lou, Q.; Bai, Y.; Fu, J.; Zhuang, C.; Zhang, Y.; et al. Robust VS4@rGO nanocomposite as a high-capacity and long-life cathode material for aqueous zinc-ion batteries. Nanoscale 2021, 13, 12370–12378. [Google Scholar] [CrossRef]
  41. Gao, S.; Ju, P.; Liu, Z.; Zhai, L.; Liu, W.; Zhang, X.; Zhou, Y.; Dong, C.; Jiang, F.; Sun, J. Electrochemically induced phase transition in a nanoflower vanadium tetrasulfide cathode for high-performance zinc-ion batteries. J. Energy Chem. 2022, 69, 356–362. [Google Scholar] [CrossRef]
  42. Zhao, C.; Yu, C.; Zhang, M.; Sun, Q.; Li, S.; Banis, M.N.; Han, X.; Dong, Q.; Yang, J.; Wang, G.; et al. Enhanced sodium storage capability enabled by super wide-interlayer-spacing MoS2 integrated on carbon fibers. Nano Energy 2017, 41, 66–74. [Google Scholar] [CrossRef]
  43. Liu, Y.; Zhao, D.; Liu, H.; Umar, A.; Wu, X. High performance hybrid supercapacitor based on hierarchical MoS2/Ni3S2 metal chalcogenide. Chin. Chem. Lett. 2019, 30, 1105–1110. [Google Scholar] [CrossRef]
  44. Liang, Y.; Feng, R.; Yang, S.; Ma, H.; Liang, J.; Chen, J. Rechargeable Mg batteries with graphene-like MoS2 cathode and ultrasmall Mg nanoparticle anode. Adv. Mater. 2011, 23, 640–643. [Google Scholar] [CrossRef]
  45. 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]
  46. Jiao, Y.; Mukhopadhyay, A.; Ma, Y.; Yang, L.; Hafez, A.M.; Zhu, H. Ion transport nanotube assembled with vertically aligned metallic MoS2 for high rate lithium-ion batteries. Adv. Energy Mater. 2018, 8, 1702779. [Google Scholar] [CrossRef]
  47. 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]
  48. Huang, M.; Mai, Y.; Fan, G.; Liang, X.; Fang, Z.; Jie, X. Toward fast zinc-ion storage of MoS2 by tunable pseudocapacitance. J. Alloys Compd. 2021, 871, 159541. [Google Scholar] [CrossRef]
  49. Sheng, Z.; Qi, P.; Lu, Y.; Liu, G.; Chen, M.; Gan, X.; Qin, Y.; Hao, K.; Tang, Y. Nitrogen-doped metallic MoS2 derived from a metal-organic framework for aqueous rechargeable zinc-ion batteries. ACS Appl. Mater. Interfaces 2021, 13, 34495–34506. [Google Scholar] [CrossRef]
  50. 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]
  51. Yao, Z.; Zhang, W.; Ren, X.; Yin, Y.; Zhao, Y.; Ren, Z.; Sun, Y.; Lei, Q.; Wang, J.; Wang, L.; et al. A volume self-regulation MoS2 superstructure cathode for stable and high mass-loaded Zn-ion storage. ACS Nano 2022, 16, 12095–12106. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, Y.; Wu, X. Strategies for constructing manganese-based oxide electrode materials for aqueous rechargeable zinc-ion batteries. Chin. Chem. Lett. 2022, 33, 1236–1244. [Google Scholar] [CrossRef]
  53. Minakshi, M.; Singh, P.; Issa, T.B.; Thurgate, S.; Marco, R.D. Lithium insertion into manganese dioxide electrode in MnO2/Zn aqueous battery. J. Power Source 2004, 138, 319–322. [Google Scholar] [CrossRef]
  54. Sun, W.; Wang, F.; Hou, S.; Yang, C.; Fan, X.; Ma, Z.; Gao, T.; Han, F.; Hu, R.; Zhu, M.; et al. Zn/MnO2 Battery Chemistry With H+ and Zn2+ Coinsertion. J. Am. Chem. Soc. 2017, 139, 9775–9778. [Google Scholar] [CrossRef] [PubMed]
  55. Chao, D.; Zhou, W.; Ye, C.; Zhang, Q.; Chen, Y.; Gu, L.; Davey, K.; Qiao, S.Z. An electrolytic Zn-MnO2 battery for high-voltage and scalable energy storage. Angew. Chem. Int. Ed. 2019, 58, 7823–7828. [Google Scholar] [CrossRef] [PubMed]
  56. Chen, X.; Li, W.; Xu, Y.; Zeng, Z.; Tian, H.; Velayutham, M.; Shi, W.; Li, W.; Wang, C.; Reed, D.; et al. Charging activation and desulfurization of MnS unlock the active sites and electrochemical reactivity for Zn-ion batteries. Nano Energy 2020, 75, 104869. [Google Scholar] [CrossRef]
  57. Ma, S.C.; Sun, M.; Sun, B.Y.; Li, D.; Liu, W.L.; Ren, M.M.; Kong, F.G.; Wang, S.J.; Guo, Z.X. In situ preparation of manganese sulfide on reduced graphene oxide sheets as cathode for rechargeable aqueous zinc-ion battery. J. Solid State Chem. 2021, 299, 122166. [Google Scholar] [CrossRef]
  58. Jiang, K.; Zhou, Z.; Wen, X.; Weng, Q. Fabrications of high-performance planar zinc-ion microbatteries by engraved soft templates. Small 2021, 17, 2007389. [Google Scholar] [CrossRef]
  59. Lv, R.; Robinson, J.A.; Schaak, R.E.; Sun, D.; Sun, Y.; 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, 56–64. [Google Scholar] [CrossRef]
  60. 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]
  61. Yang, C.; Feng, J.; Lv, F.; Zhou, J.; Lin, C.; Wang, K.; Zhang, Y.; Yang, Y.; Wang, W.; Li, J.; et al. Metallic graphene-like VSe2 ultrathin nanosheets: Superior potassium-ion storage and their working mechanism. Adv. Mater. 2018, 30, 1800036. [Google Scholar] [CrossRef]
  62. Wu, Y.; Chen, H.; Zhang, L.; Li, Q.; Xu, M.; Bao, S.J. A rough endoplasmic reticulum-like VSe2/rGO anode for superior sodium-ion capacitors. Inorg. Chem. Front. 2019, 6, 2935–2943. [Google Scholar] [CrossRef]
  63. Ming, F.; Liang, H.; Lei, Y.; Zhang, W.; Alshareef, H.N. Solution synthesis of VSe2 nanosheets and their alkali metal ion storage performance. Nano Energy 2018, 53, 11–16. [Google Scholar] [CrossRef]
  64. 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]
  65. Cai, S.; Wu, Y.; Chen, H.; Ma, Y.; Fan, T.; Xu, M.; Bao, S.J. Why does the capacity of vanadium selenide based aqueous zinc ion batteries continue to increase during long cycles? J. Colloid Interface Sci. 2022, 615, 30–37. [Google Scholar] [CrossRef]
  66. Lin, Y.; Chen, G.; Wan, H.; Chen, F.; Liu, X.; Ma, R. 2D free-standing nitrogen-doped Ni-Ni3S2 @carbon nanoplates derived from metal-organic frameworks for enhanced oxygen evolution reaction. Small 2019, 15, 1900348. [Google Scholar] [CrossRef]
  67. Liu, Y.; Hu, P.; Liu, H.; Song, J.; Umar, A.; Wu, X. Toward a high performance asymmetric hybrid capacitor by electrode optimization. Inorg. Chem. Front. 2019, 6, 2824–2831. [Google Scholar] [CrossRef]
  68. 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]
  69. Liu, Z.; Sun, H.; Qin, L.; Cao, X.; Zhou, J.; Pan, A.; Fang, G.; Liang, S. Interlayer doping in layered vanadium oxides for low-cost energy storage: Sodium-ion batteries and aqueous zinc-ion batteries. ChemNanoMat 2020, 6, 1553–1566. [Google Scholar] [CrossRef]
  70. Tong, X.; Li, Y.; Pang, N.; Zhou, Y.; Wu, D.; Xiong, D.; Xu, S.; Wang, L.; Chu, P.K. Highly active cobalt-doped nickel sulfide porous nanocones for high-performance quasi-solid-state zinc-ion batteries. J. Energy Chem. 2022, 66, 237–249. [Google Scholar] [CrossRef]
  71. Wang, Q.; Liu, T.; Chen, Y.; Wang, Q. The Ni/Ni3S2 nanocomposite derived from Ni-ZIF with superior energy storage performance as cathodes for asymmetric supercapacitor and rechargeable aqueous zinc ion battery. J. Alloys Compd. 2022, 891, 161935. [Google Scholar] [CrossRef]
Figure 1. The classification of typical TMCs cathodes and the strategies of modification.
Figure 1. The classification of typical TMCs cathodes and the strategies of modification.
Nanomaterials 12 03298 g001
Figure 2. The electrochemical performance of VS2 electrodes. (a) Atomic structure of VS2 materials. (b) The working mechanism of cathode. Reproduced with permission [31]. Copyright 2019, Elsevier B.V. (c) Rate capability of rGO–VS2 and VS2. (d) Cycling performance of these two electrodes. Reproduced with permission [32]. Copyright 2020, Elsevier B.V. (e) Zn2+ storage mechanism. Reproduced with permission [35]. Copyright 2021, Elsevier Inc.
Figure 2. The electrochemical performance of VS2 electrodes. (a) Atomic structure of VS2 materials. (b) The working mechanism of cathode. Reproduced with permission [31]. Copyright 2019, Elsevier B.V. (c) Rate capability of rGO–VS2 and VS2. (d) Cycling performance of these two electrodes. Reproduced with permission [32]. Copyright 2020, Elsevier B.V. (e) Zn2+ storage mechanism. Reproduced with permission [35]. Copyright 2021, Elsevier Inc.
Nanomaterials 12 03298 g002
Figure 3. The energy storage mechanism and cycling performance. (a) The comparison of cycling performance at 1 A g−1. Reproduced with permission [36]. Copyright 2020 Wiley-VCH GmbH. (b) The crystal structure of VS4 materials. Reproduced with permission [37]. Copyright 2020, The Royal Society of Chemistry. (c) The charge–discharge curves at 0.25 A g−1. (d) Schematic of the energy storage mechanism of Zn–VS4/CNTs batteries. Reproduced with permission [41]. Copyright 2021, Elsevier B.V. and Science Press.
Figure 3. The energy storage mechanism and cycling performance. (a) The comparison of cycling performance at 1 A g−1. Reproduced with permission [36]. Copyright 2020 Wiley-VCH GmbH. (b) The crystal structure of VS4 materials. Reproduced with permission [37]. Copyright 2020, The Royal Society of Chemistry. (c) The charge–discharge curves at 0.25 A g−1. (d) Schematic of the energy storage mechanism of Zn–VS4/CNTs batteries. Reproduced with permission [41]. Copyright 2021, Elsevier B.V. and Science Press.
Nanomaterials 12 03298 g003
Figure 4. The modification of MoS2 materials and DFT calculation. (a) Cycling stability of Zn/E–MoS2 batteries at 1.0 A g−1. (b) Schematics of the mechanism of batteries. Reproduced with permission [45]. Copyright 2018, Elsevier B.V. (c) Schematic of the MoS2–CTAB superlattice nanosheets and the d-spacing of the samples. (d) The charge density of pristine MoS2 and MoS2–CTAB nanosheets during the insertion of Zn2+. Reproduced with permission [51]. Copyright 2022, American Chemical Society.
Figure 4. The modification of MoS2 materials and DFT calculation. (a) Cycling stability of Zn/E–MoS2 batteries at 1.0 A g−1. (b) Schematics of the mechanism of batteries. Reproduced with permission [45]. Copyright 2018, Elsevier B.V. (c) Schematic of the MoS2–CTAB superlattice nanosheets and the d-spacing of the samples. (d) The charge density of pristine MoS2 and MoS2–CTAB nanosheets during the insertion of Zn2+. Reproduced with permission [51]. Copyright 2022, American Chemical Society.
Nanomaterials 12 03298 g004
Figure 5. The long-term cycle performance of MnS electrodes. (a) Rate capability of the two electrode. (b) The long-term cycling at 10 C. Reproduced with permission [56]. Copyright 2020, Elsevier Ltd. (c) Crystal structure model of α-MnS; (d) Cyclic stability in 2 M ZnSO4 and 0.1 M MnSO4 aqueous electrolyte. Reproduced with permission [58]. Copyright 2021, Wiley-VCH GmbH.
Figure 5. The long-term cycle performance of MnS electrodes. (a) Rate capability of the two electrode. (b) The long-term cycling at 10 C. Reproduced with permission [56]. Copyright 2020, Elsevier Ltd. (c) Crystal structure model of α-MnS; (d) Cyclic stability in 2 M ZnSO4 and 0.1 M MnSO4 aqueous electrolyte. Reproduced with permission [58]. Copyright 2021, Wiley-VCH GmbH.
Nanomaterials 12 03298 g005
Figure 6. The kinetic process and cycle capability of a VSe2-based Zn battery. (a) The schematic of charge density map after zinc-ion intercalation. (b) The optimal diffusion pathway of zinc ions. (c) Schematic of the two-step Zn2+ intercalation/de-intercalation process. Reproduced with permission [64]. Copyright 2020, Wiley-VCH GmbH. (d) Long-cycles performance at 5 A g−1. Reproduced with permission [65]. Copyright 2022, Elsevier Inc.
Figure 6. The kinetic process and cycle capability of a VSe2-based Zn battery. (a) The schematic of charge density map after zinc-ion intercalation. (b) The optimal diffusion pathway of zinc ions. (c) Schematic of the two-step Zn2+ intercalation/de-intercalation process. Reproduced with permission [64]. Copyright 2020, Wiley-VCH GmbH. (d) Long-cycles performance at 5 A g−1. Reproduced with permission [65]. Copyright 2022, Elsevier Inc.
Nanomaterials 12 03298 g006
Figure 7. The electrochemical performance of Zn/Ni3S2 batteries. (a) Schematic of nickel sulfide-adsorbing hydroxide radicals and being oxidized to form nickel hydroxide and sulfate radicals. (b) Gibbs free energies of Zn/Ni3S2 batteries. (c) The cycling stability of the flexible ZC12NS battery. Reproduced with permission [70]. Copyright 2021, Elsevier B.V. and Science Press. (d) Rate capability. (e) The long-term performance. Reproduced with permission [71]. Copyright 2021, Elsevier B.V.
Figure 7. The electrochemical performance of Zn/Ni3S2 batteries. (a) Schematic of nickel sulfide-adsorbing hydroxide radicals and being oxidized to form nickel hydroxide and sulfate radicals. (b) Gibbs free energies of Zn/Ni3S2 batteries. (c) The cycling stability of the flexible ZC12NS battery. Reproduced with permission [70]. Copyright 2021, Elsevier B.V. and Science Press. (d) Rate capability. (e) The long-term performance. Reproduced with permission [71]. Copyright 2021, Elsevier B.V.
Nanomaterials 12 03298 g007
Figure 8. The future target of TMCs cathodes for AZIBs.
Figure 8. The future target of TMCs cathodes for AZIBs.
Nanomaterials 12 03298 g008
Table 1. The electrochemical performance of TMC cathodes.
Table 1. The electrochemical performance of TMC cathodes.
CathodesAnodesVoltage (V)ElectrolyteCapacity (mAh g−1)Cycle Stability
(mAh g−1, Cycles, A g−1)
Ref.
VS2@VOOHZinc foil0.4–1.0 3 M ZnSO4165 (0.1 A g−1)91.4, 400, 2.5[31]
VS2 and N-doped carbonZinc foil0.2–1.8 3 M Zn(CF3SO3)2203 (0.05 A g−1)144, 600, 1[33]
VS2/VOxZinc foil0.1–1.825 M ZnCl2260 (0.1 A g−1)75%, 3000, 1[36]
VS4Zinc foil0.2–1.61 M ZnSO4310 (0.1 A g−1)110, 500, 2.5[37]
VS4/V2O3Zinc foil0.3–1.23 M Zn(CF3SO3)2163 (0.1 A g−1)-[38]
VS4@rGOZinc foil0–1.81 M Zn(CF3SO3)2450 (0.5 A g−1)82%, 3500, 10[40]
MoS2Deposited zinc on carbon cloth0.5–1.52 M ZnSO4202.6 (0.1 A g−1)164.5, 600, 1[45]
MoS2-CTABZn anode plated on carbon paper0.2–1.33 M ZnSO4181.8 (0.1 A g−1)92.8%, 2100, 10[51]
MnSzinc powder0.9–1.952 M ZnSO4 and 0.1 M MnSO4297 (0.1 A g−1)150, 1000, 1[58]
VSe2Zinc foil0.2–1.62 M ZnSO4131.8 (0.1 A g−1)80.8%, 500, 0.1[64]
Ni/Ni3S2Zinc foil1.3–1.96 M KOH and 0.6 M ZnO220 (0.2 A g−1)93.1%, 1000, 4[71]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Y.; Wu, X. Recent Advances of Transition Metal Chalcogenides as Cathode Materials for Aqueous Zinc-Ion Batteries. Nanomaterials 2022, 12, 3298. https://doi.org/10.3390/nano12193298

AMA Style

Liu Y, Wu X. Recent Advances of Transition Metal Chalcogenides as Cathode Materials for Aqueous Zinc-Ion Batteries. Nanomaterials. 2022; 12(19):3298. https://doi.org/10.3390/nano12193298

Chicago/Turabian Style

Liu, Ying, and Xiang Wu. 2022. "Recent Advances of Transition Metal Chalcogenides as Cathode Materials for Aqueous Zinc-Ion Batteries" Nanomaterials 12, no. 19: 3298. https://doi.org/10.3390/nano12193298

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