Engineered molybdenum disulfide-based nanomaterials for capacitive deionization applications

Maiyong Zhu *, Xinyue Xiang , Xinyao Jiang, Yu Yang and Kai Zhang
School of Materials Science & Engineering, Jiangsu University, Zhenjiang, 212013, P. R. China. E-mail: maiyongzhu@ujs.edu.cn

Received 8th December 2024 , Accepted 27th March 2025

First published on 27th March 2025


Abstract

Capacitive deionization (CDI) is a highly promising technique used for the removal of ions from water, showing great potential in the desalination of salty water and wastewater remediation. CDI has the advantages of energy efficiency, simple operation, excellent reversibility, long-term stability, and high feasibility for coupling with other techniques. Similar to supercapacitors, the electrode materials play a crucial role in determining the CDI performance, such as operating voltage, desalination capacity, and lifecycle of CDI cells. Molybdenum disulfide (MoS2), a typical two-dimensional (2D) metal sulfide, has gained tremendous attention in the CDI technique due to its exceptional mechanical, electrical, and optical properties. Herein, we critically discuss the inventory and recent progress in the rational design of MoS2-based electrodes for CDI cells. Initially, we present a brief introduction on the foundation knowledge of CDI systems and the structure of MoS2. For a comprehensive review, we summarize the common techniques employed to prepare MoS2-based nanomaterials, ranging from various exfoliation processes to chemical vapor deposition, colloidal synthesis, hydrothermal/solvothermal synthesis, and molten salt synthesis. Significantly, the recent progress in MoS2-based electrodes for application in CDI is summarized in detail. These systems are divided into pristine MoS2 and various MoS2-based composites with other species, such as carbon, conducting polymers, metal oxides, MXenes, and C3N4. Finally, to aid in the further development of MoS2 electrodes for efficient and long-term stable CDI, some challenges and possible solutions are outlined.


image file: d4qi03147d-p1.tif

Maiyong Zhu

Maiyong Zhu received his PhD degree from Yangzhou University (China) in 2011. In 2012, he started independent research work at school of Materials Science & Engineering, Jiangsu University (China), as an assistant professor. In 2015, he was promoted to associate professor. In 2020, he worked as a visiting professor at Kyoto University (Japan) under the support of the China Scholarship Council. Currently, the research of his group covers a wide range, including green strategies for synthesizing advanced functional materials for energy/environment related applications, valorization and recycling of solid wastes, and so on.

image file: d4qi03147d-p2.tif

Xinyue Xiang

Xinyue obtained a bachelor's degree in Inorganic Non-metallic Materials Engineering from the School of Urban Construction, Anhui Jianzhu University in June 2024. Currently, she is pursuing a master's degree at the School of Materials Science and Engineering, Jiangsu University, under the guidance of Dr. Maiyong Zhu. Her research focuses on the application of two-dimensional layered materials in capacitive deionization.

1 Introduction

Owing to the rapid industrialization and growing population, the past decades have witnessed a dramatic decrease in groundwater levels and overconsumption of fresh water, resulting in the shortage of fresh water. It is estimated that more than 4 billion people worldwide are affected by severe freshwater shortages. Thus, considering sustainable development, there are at two efficient strategies to address the issues of freshwater shortage and water resource pollution. On the one hand, water reuse and wastewater recovery demonstrate great promise in saving water resources. On the other hand, seawater or brackish water, accounting for 97% of the Earth's water volume, can be converted into fresh water through appropriate technology. Thus, considering these goals, one of the most critical tasks is to decrease the ion contents in water.1

Capacitive deionization (CDI) is a technique that removes cations and anions from water at the anode and cathode with the assistance of an external voltage, respectively. In the few past decades, as a significant process, CDI has been widely applied in wastewater remediation, resource extraction, water softening, and production of fresh water from seawater.2,3 Compared to commercial deionization techniques, such as reverse osmose, adsorption, and precipitation, CDI delivers superior advantages in terms of energy efficiency, capability to remove ions with ultralow concentration, and operation simplicity. Similar to supercapacitors, electrodes play an important role in determining the performance of CDI devices. In recent years, transition metal dichalcogenides (TMDs), with the common formula of MX2 (M represents a transition metal and X represents S, Se, or Te), have demonstrated great promise in CDI owing to their unique semiconductor features, such as narrow band gap and outstanding carrier mobility.4–6 Based on their structure, there are three representative crystal phases for MoX2, including 1T, 2H, and 3R phases. As presented in Fig. 1, MX2 typically possesses a layer structure. The interlayer interaction has been evidenced to be van der Waals not covalent bonds, whereas the intralayer interaction is M–X bond, either covalent or ionic, as determined by the electronegative difference between M and X. Among these structures, the 1T phase is thermodynamically metastable, exhibiting metallic behavior. The 2H phase, which is comprised of two layers of X/Mo/X sheets in a Bernal stacking arrangement, is thermodynamically stable, possessing semiconductor characteristics. The 3R phase is the lowest energy metastable phase and is described by a rhombohedral unit cell. It is common to describe and visualize the unit cell with hexagonal axes and these phases can be transformed among each other under certain conditions. This diversity of structural features endows MX2 with many fascinating properties, resulting in its wide application in many areas.7–12


image file: d4qi03147d-f1.tif
Fig. 1 Different phases of MX2: 1T (tetragonal symmetry, one layer per repeating unit), 2H (hexagonal symmetry, two layers per repeating unit), and 3R (rhombohedral symmetry, three layers per repeating unit). Reproduced from ref. 13 with permission. Copyright 2024, published by the American Chemical Society under a CC-BY license.

Molybdenum disulfide (MoS2), as a typical TMD, has gained increasing attention due to its ease preparation and low cost. During the past few years, a great deal of research on bare, layered MoS2 CDI materials has been conducted due to their layered nature. However, considering the intrinsically poor electrical conductivity, low removal efficiency (capacity/rate), and poor selectivity toward target ions of MoS2, scientists and engineers made many attempts to improve its CDI performance, including expanding its interlayer spacing and boosting its electrochemical activity. Some strategies, including introducing defects, intercalating guest species, doping with foreign atoms, forming heterostructures, and hybridization with other materials, have been used to improve the performance of MoS2 CDI cells. Fig. 2 provides the timeline of MoS2-based materials applied in CDI, and also outlines the key progress achieved in the past years in the development of MoS2-based CDI devices.


image file: d4qi03147d-f2.tif
Fig. 2 Timeline illustrating the key progress in MoS2-based CDI devices achieved during the past few years. Reproduced from ref. 14 with permission. Copyright 2017, Elsevier. Reproduced from ref. 15 with permission. Copyright 2018, Elsevier. Reproduced from ref. 16 with permission. Copyright 2020, Elsevier. Reproduced from ref. 17 with permission. Copyright 2022, Elsevier. Reproduced from ref. 18 with permission. Copyright 2023, Elsevier. Reproduced from ref. 19 with permission. Copyright 2025, Elsevier.

To date, some excellent reviews have been reported, either highlighting the specific synthetic methods or application of MoS2. For example, Aggarwal et al.20 provided a comprehensive review focusing on few-layer exfoliated MoS2 for application in various fields, including biosensing, gas sensing, catalysis, photodetectors, energy storage devices, light-emitting diodes (LEDs), and adsorption. Jia et al.21 summarized the recent advances in doping strategies to improve the electrocatalytic performance of MoS2 toward the hydrogen evolution reaction. Dong et al.22 highlighted the synthetic strategies for MoS2, focusing on their application in agricultural contaminant control. Yu et al.23 provided a review on the applications of MoS2 for modulating reactive oxygen species, highlighting its catalytic activity, photoluminescent behavior, and piezoelectric properties. Wang et al.24 summarized the application of MoS2 nanosheets in environmental remediation, such as adsorptive/photocatalytic elimination of contaminants, membrane isolation, sensing, and disinfection. Singh et al.25 summarized the different strategies to generate large-area MoS2 atomically thin layers for flexible electronic applications. Li et al.26 summarized the methods for the preparation of low-dimensional MoS2 nanostructures toward microwave applications.

However, despite these great achievements, a timely comprehensive review on MoS2-based materials for CDI application is obviously lacking. Thus, based on this context, herein we provide the necessary review. Initially, the theory of CDI and structural characteristics of MoS2 are briefly discussed, followed by a systematic overview on the common methods to prepare MoS2-based nanomaterials. For convenience, these methods are classified according to their operation principles, including exfoliation, chemical vapor deposition, colloidal synthesis, hydrothermal/solvothermal, and salt-assisted synthesis. More importantly, the application of MoS2-based CDI materials is summarized including pristine MoS2 and various composites. Finally, we outline the challenges and bottlenecks in the future development of MoS2-based CDI devices. It can be found that MoS2 nanostructures have illuminated boundless roadmaps for further advancement as high-performance CDI materials.

2 Synthesis methods

Generally, there are two ways to prepare molybdenum disulfide nanosheets, i.e., bottom-up and top-down methods. The former method refers to various exfoliation processes using bulk MoS2 as the precursor. To date, great achievement has been gained in exfoliating bulk MoS2. The latter method refers to the various preparation processes in which the precursors of molybdenum and sulfur undergo reactions to generate MoS2 nanomaterials. The common top-down methods include chemical vapor deposition, colloidal synthesis, hydrothermal/solvothermal synthesis, and salt-assisted synthesis. In this section, we systematically summarize these synthetic methods. It should be noted that many cases in this section may focus on thin MoS2 (either monolayer or few-layer) nanosheets owing to their wide applications given that they possess a bandgap of 1.8 eV, which is much wider than that of bulk MoS2.

2.1 Exfoliation

Similar to many other transition metal dichalcogenides, natural MoS2 exists in bulk. As a result, it is necessary to exfoliate it into single or few-layered nanosheets for optimizing its application performances. The conversion of bulk MoS2 to mono- and few-layer MoS2 can be done via the process of exfoliation in organic solvents and aqueous media. In this section, we briefly describe the process for the exfoliation of MoS2. To date, many exfoliation techniques have been developed to obtain thickness-controlled MoS2 nanosheets. Among the current reported approaches, liquid-phase exfoliation under sonication irradiation has been demonstrated to be the most versatile given that it produces distinct advantages, such as ease of processing, simplicity, and versatility. Nguyen et al.27 developed a general exfoliation process to prepare TMD nanosheets with a size of approximately a few hundred nanometers. As presented in Fig. 3a, the exfoliation is carried out in an ultrasonicator. This sonication process has been demonstrated to be feasible for preparing a variety of TMDs from the corresponding TMDs, such as MoS2, WS2, TaS2, and TiS2. Fig. 3b–e show the SEM images of the obtained TMDs, which are approximately 100–500 nm in size. AFM observation revealed that the thickness of these TMDs is about 3–4 nm. Owing to this small size, many edge sites can be obtained, endowing these TMD nanosheets with an excellent hydrogen evolution performance. Considering the long time required for the sonication process, Mohiuddin et al.28 demonstrated that the introduction of an electric field can offer new opportunities for fast and efficient exfoliation processes. The efficiency improvement can be ascribed to the combination of shear force and an electric field, making the exfoliation process much more efficient. In the sonication exfoliation process, the properties of the utilized solvent have a significant impact on the resulting 2D material, such as boiling point, surface tension, and energy, as well as solubility parameters. Nguyen et al.29 developed a two-solvent grinding-assisted sonication exfoliation process to prepare MoS2 nanosheets with single- and few-layer thick flakes. They found that the quality of the resulting MoS2 nanosheets and flake dimensions were dependent on the grinding solvent. Gupta et al.30 reported that the addition of trace water to NMP is beneficial for stabilizing MoS2 nanosheets in NMP dispersions. Their controlled investigation revealed that the MoS2 nanosheets are fragmented and chemically unstable in the absence of water. Chavalekvirat et al.31 reported an exfoliation process using water and NMP to produce MoS2 nanosheets. They found that the utilization of a mixed solvent resulted in a remarkable exfoliation yield. Under the optimal conditions, the yield of MoS2 nanosheets reached up to ca. 1.26 mg mL−1. Nevertheless, higher contents of water are not always a better case given that the system may generate MoO3 when the content of water exceeds 66.7% v/v.
image file: d4qi03147d-f3.tif
Fig. 3 (a) Schematic of the formation of TMD nanosheets via liquid exfoliation with ultrasonication. SEM images of the obtained (b) MoS2, (c) WS2, (d) TaS2, and (e) TiS2 nanosheets. Reproduced from ref. 27 with permission. Copyright 2016, the American Chemical Society. (f) Nafion-assisted exfoliation of MoS2. (g) Digital images of N-MoS2 dispersion with chemical structures of MoS2 and Nafion, respectively. (h and i) Nx-MoS2 (x = volume of Nafion in mL) dispersion in water before (h) and after (i) centrifugation. (j) Observation of the Tyndall effect from dispersed MoS2 with varying concentrations of Nafion. Reproduced from ref. 39 with permission. Copyright 2018, the American Chemical Society.

It was noted that the MoS2 nanosheets directly exfoliated in aqueous solution readily aggregated owing to the high surface tension of water. Thus, to solve this problem, surfactants or polymers are usually applied, which are inserted between the layers of MoS2 nanosheets.32 For example, Bang et al.33 obtained single-layer MoS2 sheets in NMP with NaOH assistant. They found that the addition of NaOH improved the exfoliation efficiency and dispersibility of the MoS2 nanosheets. Similarly, Li et al.34 reported that the concentration of MoS2 nanosheets can reach 1.44 mg mL−1 with the highest yield of 4.8% by adding sodium citrate to NMP. This concentration and yield are much higher than that obtained by exfoliation carried out in pure NMP under the optimized conditions, i.e., 0.96 mg mL−1 and 3.2%, respectively. Yu et al.35 also improved the exfoliation efficiency and restacking of nanosheets by using sonopolymer surfactant in 1,2-dichlorobenzene as the solvent. The improvement can be ascribed to the functionalization effect of the used sonopolymer. Owing to the expanded spacing provided by the sonopolymer, the attraction between layers dramatically decreased compared with the bare MoS2, enabling their facile redispersion, which was verified by UV–vis spectroscopy. Varrla et al.36 demonstrate the shear-exfoliation of MoS2 nanosheets in aqueous surfactant solution using a kitchen blender. Their result indicated that the mixing parameters, including the MoS2 concentration, Ci; mixing time, t; liquid volume, V; and rotor speed, N, have a significant impact on the nanosheet concentration and production rates. Under the optimal conditions, a high nanosheet concentration and production rate were achieved, reaching up to 0.4 mg mL−1 and 1.3 mg min−1, respectively. With the assistance of the cationic surfactant cetyltrimethylammonium bromide (CTAB), Gupta et al.37 prepared MoS2 nanosheets via sonication exfoliation in aqueous solution, demonstrating excellent stability when dispersed in aqueous solution. The excellent stability can be ascribed to the interaction between MoS2 nanosheets and the CTAB surfactant chains. The same group38 also investigated the surface chemistry and the role of surfactant bonding in sonication-exfoliated MoS2 nanosheets by controlling the sign of the charge on the MoS2 nanosheets, either positive or negative. The solution NMR investigation indicated that the surfactant chains on the surface of MoS2 undergo rapid exchange given that the bonding interaction between the surfactant and MoS2 are weak, which was also evidenced by in situ nuclear Overhauser effect spectroscopy.

Considering that the covalent/noncovalent interaction between MoS2 and surfactant deteriorates the properties of the starting materials and the rinsing with surfactant may worsen the exfoliated flakes, leading to restacking, Oh et al.39 proposed the Nafion-assisted exfoliation of MoS2 (Fig. 3f). The exfoliation driving force can be ascribed to the hydrophobic interaction between the Nafion molecules and MoS2 layer, as presented in Fig. 3g. Owing to simultaneous presence of both hydrophilic and hydrophobic moieties, the Nafion-assisted exfoliated MoS2 showed good dispersibility. As shown in Fig. 3h and i, the dispersibility of the exfoliated MoS2 demonstrated little dependence on the variation in the amount of Nafion, evidencing the positive influence of Nafion surfactant on the dispersion efficiency. Furthermore, the improved dispersibility of all the Nafion-assisted exfoliated MoS2 was also confirmed by the Tyndall effect, as presented in Fig. 3j. Karunakaran et al.40 employed thiol-containing surfactant-assisted exfoliation to generate functionalized 2H-MoS2. The resulting functionalized 2H-MoS2 exhibited highly enhanced antibacterial efficiency. Melaku et al.41 developed a programmed exfoliation process to generate water-dispersible exfoliated MoS2 nanosheets, possessing well-tunable structural characteristics using an environmentally friendly water-soluble cytosine-functionalized supramolecular polymer (Cy-PPG). In their work, dozen-layered MoS2 nanosheets could be obtained by exfoliating MoS2 under ultrasound irradiation.

In addition, several biomacromolecules (such as polysaccharides, pullulan, cyclodextrin, wool keratin, natural polyphenols, and amino acid) have also been reported to assist in the sonication exfoliation of bulk MoS2 and dispersing exfoliated MoS2 nanosheets in water.42–47 For example, Sonker et al.48 presented a one-step exfoliation process to obtain water-stable MoS2 nanosheet suspensions using sulfated cellulose nanocrystals (CNC) as surfactants and exfoliating agents. A high extraction efficiency of 34 wt% was achieved, which is much higher than that using CNF as the exfoliating agent (21 wt%).49 Yadav et al.50 developed an ultrasonication process in a bio-based polymer-solvent to prepare quasi 2D-TMDs (including MoS2, TiS2, WS2, NbS2 and MoSe2) with 1–5 layers. As shown in Fig. 4, the whole exfoliation includes three steps. Initially, bulk TMD and bio-based polymer are mixed homogeneously, which is achieved by grinding. Afterward, solvent mixing and intercalation will occur by adding the mixture of TMD/polymer to the desired solvent. Finally, ultrasonication is applied to induce the exfoliation of TMD. A series of solvent-polymer systems was screened in their work to optimize the yield. The results indicated that MoS2/HPC/BuOH (I), MoS2/CMC/BuOH (II), MoS2/HPMC/BuOH (III), MoS2/HPC/IBA (IV) and MoS2/HPC/EG (V) delivered the highest yield of the corresponding 2D TMD nanosheets, respectively.


image file: d4qi03147d-f4.tif
Fig. 4 Bio-based polymer-assisted ultrasonicated exfoliation of TMD in organic solvent. Reproduced from ref. 50 with permission. Copyright 2022, Elsevier.

Recently, ionic liquids (ILs) have attracted extensive attention owing to their superior advantages, such as negligible vapor pressure, versatile combability with many precursors, and wide redox potential.51–57 ILs can serve as the reaction medium, electrolyte, catalyst, structure directing agent, etc. Tian et al.58 successfully exfoliated bulk MoS2 into MoS2 ultrathin nanosheets via sonication in water with the assistance of IL crystal (ILC), which served as both the exfoliation and dispersing agent (Fig. 5a). The driving force for the exfoliation can be ascribed to the cation–π interaction between the imidazolium in ILC and MoS2. Meanwhile, this interaction can prevent the re-aggregation of the exfoliated MoS2 nanosheets. As presented in Fig. 5b, the resulting MoS2 nanosheets showed poor dispersibility in water before adding ILCs. Once a small amount of ILCs was introduced, they could be homogeneously dispersed. Furthermore, high aggregation was observed in the case of the bulk MoS2 powder before adding ILCs, which is ascribed to their strong hydrophobicity given that they possess a large contact angle (125°) (Fig. 5c). This value is much larger than that of the ILC300–MoS2 nanohybrid (7.6°). In addition, the ILC-assisted ultrasonic treatment also resulted in an increase in the concentration of ILC–MoS2 dispersion collected by centrifugation, which was determined to be approximately 7.32 mg mL−1 (Fig. 5d). The resulting ILC–MoS2 dispersion could yield a macroscopic aerogel of MoS2 after freeze-drying. The mass of the macroscopic aerogel reached about 0.43 mg. This result demonstrates the high concentration and dispersibility of the as-prepared ILC–MoS2 dispersions (Fig. 5e). Fig. 5f and g display the SEM/TEM images of the bulk MoS2 and exfoliated MoS2 nanosheets, whose lateral size were determined to be in the range of several micrometers and 80–320 nm, respectively. Guo et al.59 demonstrated that IL can serve as a gate to inject H+ into layered MoS2 to manipulate its carrier concentration. Interestingly, the injected H+ was found to be located in the interlayer of MoS2 but not the crystal lattice. Consequently, the injection of H+ had an influence on the multilayer MoS2. The injection concentration was found to be as high as ∼1.08 × 1013 cm−2. Wang et al.60 also demonstrated the IL gating of suspended few-layer MoS2 transistors. In their work, the accumulation of ions could occur on both exposed surfaces. Upon applying IL, a significant improvement in conductance was obtained for all the free-standing samples, whereas this was not the case for the substrate-supported devices. Impressively, a transition from metal to insulator could be achieved in the suspended MoS2 devices by modulating the IL gate voltage. By measuring the charge density of states in IL-gated MoS2, Lee et al.61 concluded the coexistence of monolayer MoS2 nanosheets with multilayer MoS2 in the channel area of IL-gated MoS2 transistors.


image file: d4qi03147d-f5.tif
Fig. 5 (a) Procedure for the exfoliation of bulk MoS2 into nanosheets by ILC-assisted exfoliation. (b) ILC-dependent dispersibility of bulk MoS2 in water. (c) Water contact angle of bulk MoS2 and ILC–MoS2. Photographs of ILC–MoS2 (d) dispersion and generated (e) aerogel. (f) SEM image of bulk MoS2. (g) TEM image of ILC–MoS2. Reproduced from ref. 58 with permission. Copyright 2022, the American Chemical Society.

As analogues of ILs, deep eutectic solvents (DESs), which are comprised of hydrogen bond donors and hydrogen bond acceptors, have also attracted significant attention owing to their easy preparation, excellent electronic conductivity, and negligible vapor pressure. DESs are regarded as reaction media or modification agents to generate functional materials or extract functional species from biomass.62–67 Owing to their adjustable composition and acidity, DESs show great capability to dissolve a wide range of species, showing great promising in pretreating biomass and extracting valuable species from biomass or sludge.68–73 Inspired by their excellent dissolubility, DESs can also be utilized to synthesize inorganic materials. For example, Mohammadpour et al.74 introduced a mechanical exfoliation approach to generate two-dimensional nanosheets from bulk MoS2 sugar-based natural deep eutectic solvents, which played dual roles, i.e., the exfoliation medium and intercalating solvent. A high yield of 44% was achieved using this approach. Interestingly, the thickness, length, concentration of nanosheets, and phase ratio of 2H to 1T could be conveniently tuned by adjusting the experimental parameters. Under the optimized conditions, the average thickness and length of the exfoliated nanosheets were determined to be 4 and 150 nm, respectively, as evidenced by their UV–vis absorption spectrum. The resulting nanosheets were confirmed to be comprised of 2H and 1T phases with the ratio of 2H/1T = 1.4.

Considering the liquid nature of ILs, recently poly(ionic liquid)s (PILs) have attracted extensive attention as alternatives to ILs to modify materials, generating new desired functions. Biswas et al.75 demonstrated a quick sonication process, which showed high efficiency for the preparation of single or few-layer MoS2 nanosheet dispersions in the presence of cationic PILs.76 Owing to the versatile solubility of PILs, this exfoliation can be carried out in both aqueous and organic media. Owing to the adsorption of the PIL onto the surface of the MoS2 nanosheets, the resulting PIL-stabilized nanosheet dispersions demonstrated excellent stability in both water-soluble poly(vinyl alcohol) and nonaqueous-soluble poly(methyl methacrylate) matrices. Interestingly, the PIL made the resulting MoS2 nanosheets responsive toward ions and temperature in aqueous medium. Wang et al.77 reported that thermoresponsive PILs can be used as versatile surfactants to assist the exfoliation of various 2D materials, such as MoS2, graphite, and hexagonal boron nitride, through consecutive sonication. They also demonstrated that the reliable interaction between 2D materials and the thermoresponsive PIL would facilitate exfoliation, achieving the noncovalent functionalization of the exfoliated nanosheets. Meanwhile, the resulting exfoliated nanosheet suspensions were stable, which could be reversibly tuned by temperature owing to the thermoresponsive phase transition behavior of the thermoresponsive PIL.

Gopalakrishnan et al.78 reported a liquid exfoliation technique to deliver a highly dispersed suspension of MoS2 quantum dots interspersed in few-layered MoS2 nanosheets. As presented in Fig. 6, the process involved a combination of bath sonication and ultrasound probe sonication of MoS2 flakes. The role of bath sonication is to induce the transformation of MoS2 flakes into MoS2 quantum dots. As shown in Fig. 6a, the size of the MoS2 quantum dots was about 2 nm. The driving force for this conversion can be ascribed to the hydrodynamic forces caused by the increased pressure and temperature, which resulted in the breakdown of the bulk MoS2 into quantum dots. Fig. 6b displays the TEM image of MoS2 quantum dots interspersed in few-layered MoS2 nanosheets. Benefiting from this unique structure, the resulting MoS2 nanostructures possessed a larger concentration of active edges, leading to excellent activity for the hydrogen evolution reaction. As presented in Fig. 6c and d, the onset potential and Tafel slope of the resultant MoS2 electrode were both determined to be very low, respectively. By simply controlling the duration of ultrasonication, Dai et al.79 developed a sulfuric acid-assisted ultrasonication route to synthesize various MoS2 structures including monolayer MoS2 flakes, nanoporous MoS2, and MoS2 quantum dots. In another study, Xu et al.80 reported a process of sonication followed by hydrothermal treatment to prepare MoS2/WS2 quantum dots (Fig. 6e). To improve the stability of the MoS2 nanosheets obtained by sonication-assisted exfoliation, chemical modification demonstrates versatility in modifying/adding new properties to help realize this.81 Accordingly, Vera-Hidalgo et al.82 presented a sonication-assisted exfoliation process to modify 2H-MoS2 with maleimide derivatives. In their work, the modification was achieved by a thiol–ene “click” reaction. Li et al.83 reported a sonication–hydrothermal method to prepare a CD@2D MoS2 heterostructure, regulating the energy level configuration and visible light absorption performance.


image file: d4qi03147d-f6.tif
Fig. 6 Schematic of the procedure for the synthesis of MoS2 quantum dots interspersed in MoS2 nanosheets. TEM images of (a) MoS2 quantum dots and (b) MoS2 quantum dots interspersed in the exfoliated MoS2 nanosheets. Reproduced from ref. 78 with permission. Copyright 2014, the American Chemical Society. (e) Sonication-assisted exfoliation process to synthesize MoS2/WS2 quantum dots. Reproduced from ref. 80 with permission. Copyright 2015, Wiley-VCH.

It has been noted that exfoliation under sonication conditions usually gives MoS2 with few- and several-layer thick nanosheets, making it difficult to obtain single-layer MoS2. Meanwhile, it is difficult to control the degree of exfoliation. In some cases, excessive exfoliation may lead to the formation of quantum dots. In addition, the yield of MoS2 nanosheets through this technique is not high. Thus, to overcome these issues, introducing intercalation species may be an alternative. Knirsch et al.84 demonstrated a functionalization route to bond organic groups to the surface of MoS2. As presented in Fig. 7a, n-butyllithium (n-BuLi) is initially used to react with bulk MoS2, achieving 2D nanomaterials. Owing to this intercalation, the exfoliated MoS2 will be negatively charged. Subsequently, a mild bath-type sonication is applied to the negatively charged MoS2, resulting in efficient exfoliation into individual sheets. Upon the addition of 4-methoxyphenyldiazonium tetrafluoroborate, the negative charges on MoS2 can be quenched, thus yielding functionalized MoS2 (f-MoS2). The interaction between MoS2 and 4-methoxyphenyldiazonium tetrafluoroborate was evidenced to be C–S covalent bonds. Interestingly, excessive sonication exposure may lead to the defunctionalization of f-MoS2, generating df-MoS2. Fig. 7b–d show the HRTEM images of CE-MoS2, f-MoS2, and df-MoS2, respectively. The lattice is intact over wide regions, indicating that neither functionalization nor defunctionalization caused severe structural damage to MoS2. The interlayer spacing, which could be adjusted by inserting guest species, had a significant impact on the properties of MoS2. The redox potential of organolithium is significant for the intercalation and exfoliation of bulk MoS2. Owing to the pyrophoric nature of n-BuLi, the intercalation and exfoliation process using n-LiBu is usually carried out under oxygen-free and water-free conditions to avoid a serious hazard for safe operation. Thus, to circumvent this problem, Loh et al.85 developed an alkaline metal (Li, Na, and K) naphthalenide salt-assisted expansion and intercalation process to generate MoS2. In this way, high-quality single-layer MoS2 sheets could be obtained with an unprecedentedly large flake size (up to 400 μm2). Zhu et al.86 compared several lithium reagents, including n-butyllithium (n-BuLi), naphthalenide lithium (Nap-Li) and pyrene lithium (Py-Li). Their results indicated that the redox potential of the lithium-containing compounds has significant influence. A too high potential (above 1.13 V vs. Li+/Li) may lead to the decomposition of bulk MoS2. In contrast, a too low potential (below 0.55 V vs. Li+/Li) is not capable of achieving intercalation. Fortunately, the redox potential of Py-Li was determined to be 0.86 V (vs. Li+/Li), which is suitable for precise Li+ intercalation. Interestingly, the intercalation led to the formation of the LiMoS2 lamellar compound without undesirable structural damage. Pondick et al.87 observed a 2H-1T′ phase transition induced by lithium intercalation in MoS2 nanosheets. Zhang et al.88 developed a surfactant intercalation technique to yield single-layer TMDs. As shown in Fig. 7e, C15H34ClN molecules are intercalated into the interlayer of bulk TMDs, which is a thermodynamically controlled process. This is based on the expansion of the long-chain molecule and weak van der Waals interaction between the stacking layers. By intercalating C15H34ClN molecules, the interlayer spacing demonstrated a dramatic enlargement. Owing to the generated shear force as well as the extended free spacing, TMDs can be further exfoliated to generate stable TMD suspensions in NMP with mild ultrasound. Fan et al.89 developed a fast sonication-assisted lithium intercalation route to yield exfoliated 2H MoS2. As presented in Fig. 7f, an intermediate, stoichiometric LiMoS2, is generated during the exfoliation process through the reaction of n-butyllithium (n-BuLi) for the complete intercalation of MoS2. Furthermore, hexane as the solvent serves to accelerate the intercalation. Interestingly, the resulting LiMoS2 could be further converted into 1T MoS2 by further sonication. Meanwhile, a phase transition from 2H to 1T could be achieved during the intercalation process, which was reversed by IR laser irradiation using a DVD optical drive. The same group90 later obtained MoS2 nanosheets by exfoliating crystalline MoS2, with the retention of the semiconducting 2H phase. The controlled reaction of MoS2 with substoichiometric amounts of n-BuLi led to intercalation of the edges of the crystals. Upon further exfoliation in a 45 vol% ethanol–water solution, trilayer MoS2 nanosheets were obtained, which demonstrated excellent suspension behavior.


image file: d4qi03147d-f7.tif
Fig. 7 (a) Schematic of the intercalation exfoliation of MoS2 using an organic lithium agent. HRTEM images of MoS2 obtained by different processes: (b) chemical exfoliation, (c) functionalized MoS2, and (d) defunctionalized MoS2. Reproduced from ref. 84 with permission. Copyright 2015, the American Chemical Society. (e) Schematic of the intercalation and exfoliation of TMDs using dodecyl trimethyl ammonium chloride (C15H34ClN). Reproduced from ref. 88 with permission. Copyright 2022, the American Chemical Society. (f) Reaction scheme for the preparation of exfoliated 2H MoS2. Reproduced from ref. 89 with permission. Copyright 2015, the American Chemical Society.

Liu et al.91 proposed an electrochemical exfoliation process to prepare monolayer and few-layer MoS2 nanosheets. As depicted in Fig. 8a, the bulk MoS2 was used as the cathode and Pt wire served as the anode in the system. The electrolyte was Na2SO4 solution. Under direct current assistance, the bulk MoS2 crystals could be exfoliated into monolayer or single-layer MoS2 nanosheets. Owing to its hydrophilicity, the resulting MoS2 nanosheet showed poor wettability toward water and polar organic solvents. As a result, the MoS2 nanosheet could not be well dispersed in Na2SO4 solution and NMP solvent, as shown in Fig. 8c and d, respectively. Fig. 8e illustrates the exfoliation mechanism, which can be divided into two steps. Initially, ˙OH and ˙O radicals are generated by the oxidation of water upon the application of a positive bias to the working electrode. The generated radicals as well as SO42− anions are easily assembled around the bulk MoS2 crystal by inserting themselves between the MoS2 layers. Consequently, the van der Waals interactions between the layers will be weakened. Afterward, the reactivity of the radicals and/or anions results in the release of O2 and/or SO2, expanding the interlayer spacing of the MoS2 sheets. Finally, with the assistance of the erupting gas, the MoS2 flakes are detached from the bulk MoS2 crystal, which are then suspended in the solution. It should be noted that during the electrochemical exfoliation process, oxidation reactions should occur on the surface of the bulk MoS2 crystal, which is crucial in determining the quality and degree of oxidation of the exfoliated MoS2 nanosheets. Fig. 8f–i display the TEM/HRTEM images of the resulting thin MoS2 nanosheet, which possessed a lateral size of 10 μm and a folded edge of a monolayer MoS2 nanosheet.


image file: d4qi03147d-f8.tif
Fig. 8 (a) Schematic of the exfoliation of bulk MoS2 crystal via electrochemical process. (b) Photograph of a bulk MoS2 crystal. Exfoliated MoS2 flakes suspended in Na2SO4 solution (c) and dispersed in NMP solution (d). (e) Schematic for mechanism of electrochemical exfoliation of bulk MoS2 crystal. (f–i) TEM and HRTEM images of the resulting MoS2 nanosheets. Reproduced from ref. 91 with permission. Copyright 2014, the American Chemical Society.

To improve the stability and oxidation resistance of MoS2, García-Dalí et al.92 proposed a biomolecule-assisted electrolytic exfoliation method to prepare MoS2 in aqueous solution. Benefiting from the introduction of biomolecules, the resulting MoS2 nanosheets exhibited colloidal dispersion with water, oxide-free, and phase-preservation. The mechanism study indicated that the exfoliation is driven by the penetration of hydrated cations from the electrolyte. Considering the inert characteristics of 2H-MoS2, making its further functionalization difficult, the same group93 later reported an electrolytic approach to achieve 2H-MoS2 nanosheets functionalized with organic iodide-derived molecular moieties. This electrolytic exfoliation/functionalization triggered the expansion of the MoS2 crystal in an accordion-like fashion. In another work, Zhuo et al.94 demonstrated an electrochemical exfoliation process, in which the chemical functionalization of MoS2 layers was simultaneously realized. In detail, aryl diazonium salts were applied in their system, which played two roles. On the one hand, it can be covalently grafted onto 2H-MoS2, endowing 2H-MoS2 with more functions. On the other hand, its intercalation effect aids the exfoliation process, enlarging the interlayer spacing and preventing restacking.

To date, three main exfoliation mechanisms have been reported, i.e., exfoliation induced by ultrasound or shear forces, triggered by lithium intercalation making use of organolithium compounds (e.g., n-butyllithium), and electrochemical exfoliation. Ultrasound- and shear force-induced exfoliation is preferred to provide few- and several-layer-thick nanosheets of high structural quality. However, a limitation of this technique is its low exfoliation yield. By contrast, lithium intercalation approaches, which are usually conducted in organic solvents, are capable of affording single-layer nanosheets in high yields (>50 wt%). Nevertheless, their stringent conditions (i.e., in the absence of oxygen and humidity) constitute their main drawback. Furthermore, lithium intercalation often leads to a phase transition (from 2H to 1T), generating numerous defects and imperfections in the final product. In the case of electrochemical exfoliation, the redox potential has a significant impact on the exfoliation and functionalization process. Caution should be taken during the electrochemical exfoliation process given that the anodes (such as platinum) can undergo unintentional doping, which is not desirable in most cases. In addition, owing to the lack of a stable reference electrode, the applied potential is not a reliable value. It should be noted that there are some other exfoliation techniques besides the above-discussed three techniques, such as femtosecond laser exfoliation,95 microwave-assisted exfoliation,96 plasma-induced exfoliation,97 light-induced exfoliation,98 and redox-based exfoliation.99,100

2.2 Chemical vapor deposition

Chemical vapor deposition (CVD) has been demonstrated to be a successful approach for synthesizing various 2D materials for many important technological applications, such as microelectronics, superconductors, hard coatings, and smart windows.101,102 This method has several advantages, for example low-cost and powerful method and capability to tune the nucleation density, enabling the preparation both monolayer 2D materials and interesting structures such as superlattices and super-twisted spirals. In the study by Bai et al.,103 a flow of nonreactive gas (such as N2) was introduced to carry the sulfur vapor, allowing it to react with the molybdenum precursor located downstream of the furnace (Fig. 9a–d). Utilizing a similar procedure, Wang et al.104 reported the growth of monolayer MoS2 on SiO2, where the growth substrate SiO2 was placed upside down on a quartz boat directly above the Na2MoO4·2H2O solution precursor. Tarasov et al.105 synthesized MoS2 with a large area through the direct sulfurization of Mo thin films, which were generated by evaporation and deposited on SiO2.
image file: d4qi03147d-f9.tif
Fig. 9 (a) Schematic of the solid–gas process to prepare MoS2 deposited on DWCNT. (b) SEM and (c) high-magnification SEM images of MoS2 flakes, and (d) enlarged image of the marked area. Reproduced from ref. 103 with permission. Copyright 2022, published by Wiley-VCH under a CC-BY license. (e–g) Schematic of the fabrication process of MoS2 nanoribbons from MoO3 nanoribbon precursors: (e) formation of MoO3 precursor, (f) reaction to generate hybrid MoS2/MoO2 nanoribbons and (g) conversion to MoS2 nanoribbons. (h–j) SEM images of the obtained MoO3, MoS2/MoO2 hybrid and MoS2 nanoribbons. Inset in (f) MoS2 nanoribbons synthesized in one batch. Reproduced from ref. 109 with permission. Copyright 2020, Wiley-VCH.

Recently, Mouloua et al.106 reported a chemical vapor deposition process to deposit MoS2 nanostructures on a quartz substrate, exhibiting exceptional crystalline quality and a triangular-like morphology. By placing S and Se powder upstream and MoO3 downstream, Li et al.107 prepared alloy MoS2xSe2(1−x) triangular nanosheets with complete composition tunability. The resulting MoS2xSe2(1−x) triangular nanosheets exhibited tunable optical properties, which is in good agreement with the composition of the alloy nanosheets. By controlling the growth parameters, the same group also realized the continuous lateral growth of the MoS2(1−x)Se2x alloy with a graded bilayer composition. Specially, the value of x gradually increased from x = 0 (pure MoS2) to x = 0.68 from the center to the edge of the nanosheet.108 In another work, Huang et al.109 developed a chemical vapor deposition process to synthesize MoS2. As illustrated in Fig. 9e–g, their synthesis involves three steps. The first step is the formation of the MoO3 beam precursor, which is achieved through a conventional hydrothermal procedure, briefly heating a solution containing (NH4)2MoO4·4H2O and nitric acid at 180 °C for 20 h. Owing to its easy operation, this step generates MoO3 nanoribbons in large quantities.110 The second step refers to the formation of the intermediate of MoS2/MoO2. This procedure is carried out in a two-zone furnace with the sulfur powder placed upstream to generate sulfur vapor. In this step, the temperature is crucial for the formation of the MoS2/MoO2 hybrid. A high temperature may change the morphology of the nanoribbons. For example, the temperature of 650 °C may cause the collapse of the MoO3 nanoribbons given that the melting temperature of bulk MoO3 powder is about 800 °C.111 Alternatively, a low temperature cannot generate sulfur vapor efficiently. Considering these factors, in their system the temperature of 500 °C was found to be suitable. The third step is the formation of the final MoS2 nanoribbons, which can be considered further sulfurization. Thus, it is not necessary to isolate the intermediate MoS2/MoO3 hybrid. To achieve complete sulfurization within a short duration, this step can be carried out at a high temperature (900 °C) given that the melting temperature of MoS2 and MoO2 is as high as 2375 °C and 1100 °C, respectively.

To reveal the formation mechanism of MoS2 in solid–gas synthesis, Dong et al.112 performed a systematic theoretical study. Their results indicated that the edge reconstruction has a significant impact on the growth kinetics of MoS2. The reconstruction is dramatically determined by the concentrations of Mo and S in the growth environment (Fig. 10a). The substrate applied to accommodate MoS2 in the solid–gas synthesis system also has a significant impact on the morphology of the resultant MoS2. For example, Chowdhury et al.113 demonstrated a gas-phase synthesis of dimensional TMD crystals without lithography. In their work, two substrates were employed, SiO2 and Si–Px (PH3-treated Si substrates). When SiO2 was used as the substrate, triangular 2D MoS2 crystals were generated (Fig. 10b and c), which is the most common phenomenon. This was not the case for the Si–Px substrate. The grown MoS2 on Si–Px produced crystalline nanoribbons, as presented in Fig. 10d and e. This different directional growth can be explained by the preferential growth from one of the edges of a nanoscale 2D crystal seed, leading to the formation of 1D MoS2 nanoribbons. Yang et al.114 used high-Miller-index Au facets as templates to synthesize monolayer TMD ribbons (e.g., MoS2, WS2, MoSe2, WSe2, and MoSxSe2−x). To unlock the morphological evolution from triangular domains to patterned ribbons, they further investigated the growth kinetics of specific edges. Furthermore, the growth process conformed well to the one-dimensional edge epitaxial growth mode, leading to the formation uniformly aligned TMD ribbons merging into single-crystal films.


image file: d4qi03147d-f10.tif
Fig. 10 (a) Schematic showing the strategy of exploring the growth mechanism of MoS2. Reproduced from ref. 112 with permission. Copyright 2023, the American Chemical Society. Scheme depicting the low-pressure synthesis of 2D MoS2 crystals on SiO2 substrate (b) and Si–Px substrate (d). Typical 2D MoS2 crystal omni-directional lateral growth mode on SiO2 substrate (c) and Si–Px substrate (e). Reproduced from ref. 113 with permission. Copyright 2019, the authors published by Springer Nature.

Kwon et al.115 developed an improved chemical vapor deposition process to prepare MeS2 (Me = Mo or W) layers with a polycrystalline structure. The diagram of the growth setup is presented in Fig. 11a and b, where the system integrates CVD and spin-coating and contains three heating zones. The chemical principle of forming MeS2 is based on the thermolysis of (NH4)2MeS4 (Me = Mo or W) under an N2 atmosphere. Fig. 11c–e display the TEM images and corresponding EDS and EDX information of the obtained MoS2 layers. These results indicate that the resultant MoS2 layers produce mainly single-crystalline characteristics.116 Nevertheless, the HRTEM images and ring motifs in the SAED patterns indicated the presence of polycrystalline areas, where the atomic arrangement is nonperiodic. The mechanism investigation revealed that the conversion of (NH4)2MeS4 into MeS2 is a two-step process, which is dependent on the applied temperature. When the temperature is 500 °C, the (NH4)2MeS4 precursor is converted into MoSx thin layers, which will be further converted into MeS2 thin films at a temperature of 900 °C. It was noted that both steps occur in the presence of H2, the role of which can be regarded as a reducing agent, as described by the following equations:

 
(NH4)2MeS4 + H2 → 2NH3 + H2S + MeS3 below 500 °C (1)
 
MeS3 + H2 → MeS2 + H2S above 900 °C (2)


image file: d4qi03147d-f11.tif
Fig. 11 (a) Fabrication of (NH4)2MeS4 precursors. (b) Thermal CVD process to generate MoS2 and WS2. Microscopic analysis of synthesized MoS2 and WS2 layers. (c–e) Characterization of MoS2. Reproduced from ref. 115 with permission. Copyright 2015, the American Chemical Society. (f) Diagram of MOCVD setup for growth of MoS2. (g) Growth time-dependent optical images of MoS2. Reproduced from ref. 118 with permission. Copyright 2015, Springer Nature.

Wree et al.117 prepared hexagonal MoS2 layers through a metalorganic CVD (MOCVD) process, in which elemental sulfur was utilized as the co-reactant and the Mo source was a molybdenum-organic compound, 1,4-di-tert-butyl-1,4-diazabutadienyl-bis(tert-butylimido)molybdenum(VI) [Mo(NtBu)2(tBu2DAD)]. Owing to the combination of imido and chelating 1,4-diazadieneyl ligand moieties around the molybdenum metal center, a monomeric could be obtained, possessing excellent thermal characteristics relevant for vapor phase deposition applications. The structure and composition characterization revealed that the resulting MoS2 films were crystalline and stoichiometric. Benefiting from the moderate process conditions, scalability and controlled targeted composition, this MOCVD process is highly promising. In another work, Kang et al.118 reported an alternative MOCVD process to generate high-mobility 4-inch wafer-scale films of monolayer molybdenum disulphide (MoS2). Fig. 11f displays a diagram of the MOCVD growth setup. In this system, all the precursors are in gaseous form, including (C2H5)2S, W(CO)6, Mo(CO)6, and H2, which are diluted in argon as a carrier gas. The concentration of each reactant can be precisely controlled during the entire growth time by regulating the partial pressure (PX) of each reactant, X. Fig. 11g shows the optical images of the resulting MoS2 at different growth times. The results indicate no nucleation of a second layer during the formation of the first layer.

The CVD process has advantages for the large-scale production of MoS2 nanosheets at low cost. In many cases, solid MoO3 and sulfur powder are used as the precursors. In this process, the greatest disadvantage is the spatial nonuniformity, which can be ascribed to the fact that these two precursors exhibit different vapor pressures and melting temperatures. Meanwhile, the growth process should be further controlled to improve the reproducibility. The emergence of organometallic precursors provides opportunities to address these problems. Consequently, CVD will still occupy a great position in synthesizing MoS2 nanosheets. Unfortunately, these metal–organic compounds are toxic, having a serious environmental impact and corrosion to instruments. Thus, in the future, more attention should be focused on addressing issues such as the low growth rate, small domain size, high production cost, and search for environmentally friendly organometallic precursors.

2.3 Colloidal synthesis

Colloidal synthesis is another popular bottom-down strategy, which has attracted significant interest in materials synthesis because of their shape-dependent physical and chemical properties. In the synthesis of 2D materials, such as nanoflakes and nanosheets, it is possible to tune their lateral sizes and stacking confinement down to the atomic scale through this method. The most commonly applied colloidal synthesis method for the preparation of MoS2 nanoflakes is based on the reaction of molybdenum compounds (such as [Mo(CO)6], [Mo(CH3COO)2]2, and MoCl5) with various sulfur sources (such as thioacetamide (TAA) and hexamethyldisilazane (HMDS)) in the presence of a surfactant (such as oleyamine or stearic acid). For example, Altavilla et al.119 reported a colloidal process to synthesize few-layer MoS2 nanosheets, wherein the single-source precursor (NH4)2MoS4 decomposed in oleylamine solvent at 360 °C. The resulting MoS2 nanosheets were covered by a dynamic protective coating of oleylamine. Consequently, the MoS2 suspension exhibited excellent stability, avoiding aggregation and oxidation phenomena. Zechel et al.239 investigated the effect of the sulfur source in the synthesis of MoS2 nanoflakes via colloidal synthesis. In their work, [Mo(CH3COO)2]2 served as the Mo source, whereas a series of sulfur-precursors was screened, including thioacetamide (TAA), 3-mercaptopropionic acid (3-MPA), L-cysteine (L-CYS), mercaptosuccinic acid (MSA), 11-mercaptoundecanoic acid (MUA), 1-dodecanethiol (DDTH), and di-tert-butyl disulfide (DTBD). They concluded that TAA, the most commonly used S-precursor, is a known carcinogen, delivering higher toxicity than the other investigated S-precursors. Sanikop et al.120 prepared MoS2 nanosheets with control of the average number of layers (N) and defect concentration via colloidal synthesis. Frauendorf et al.121 prepared MoS2 nanosheets and nanoparticles using MoCl5 as the precursor reacting with elemental sulfur and hexamethyldisilazane in oleic acid (OA) and oleylamine (OlAm) at 320 °C. Many properties of nanomaterials are dependent on their morphology/size/shape, which can be controlled by adjusting their synthetic parameters. For example, Lambora et al.122 investigated the evolution of the morphology of colloidal MoS2 nanostructures, which were synthesized using (NH4)2MoS4 as both the Mo and S source, whereas octadecene (ODE) served as the solvent. The oleic acid and oleylamine served as the surfactant to direct the morphology of the product (Fig. 12a). Similar to other wet chemical synthetic systems, the formation of these MoS2 nanostructures also involves nucleation, growth and assembly into the final product (Fig. 12b). Different from the common cases where the morphologies are usually controlled by varying the concentration/dosage of oleic acid and oleylamine, here the reaction temperature is the determining parameter. Specially, it was noted that no precipitation occurred when the reaction temperature was below 90 °C, which can be ascribed to the high pyrolysis temperature of (NH4)2MoS4 (≈155 °C). When the reaction temperature was above 160 °C, a non-emissive black precipitate was formed, which can be attributed to the ultrafast reaction occurring at high temperature. Recently, the same group123 further prepared MoS2 quantum dots using the same reaction system at 120 °C. Interestingly, by varying the OAc/OAm (mL mL−1) concentration, the optical characteristics of MoS2 quantum dots could be easily adjusted, which can be ascribed to the different microstructures of the resulting MoS2 quantum dots. Fig. 12c–h display the HRTEM images of the MoS2 quantum dots capped with different concentration combinations of OAc and OAm. All the OAc/OAm pairs produced spherical dots. In the absence of OAc or OAm, the quantum dots with an average diameter of ≈13 nm were produced, which were larger than that of the other samples. In the case of volume fractions of 3 mL/0 mL, smaller dots with an average diameter of ≈3 nm were generated. Regarding size distribution, the volume fraction of OAc = 3 produced a narrow size distribution. Liu et al.124 prepared a series of nanosized 1T′-TMD monolayers with high phase purity via the general scalable colloidal synthetic technique. These 1T′-TMD monolayers include 1T′-MoS2, 1T′-MoSe2, 1T′-WS2, and 1T′-WSe2. Moreover, the generation mechanism investigation (combination of systematic experiments and theoretical calculations) indicated that the surfactant prevented the stacking of the layers. Particularly, the resulting 1T′-MoS2 nano-monolayers demonstrated a significant HER performance with an onset overpotential of −117 mV and a current density of 10 mA cm−2 at 149 mV (vs. the RHE). Meerbach et al.125 also reported a general colloidal synthesis strategy to prepare a series of TMDs, where in the metallic chloride and carbon disulfide were applied as the metal and sulfur source, respectively. Besides a single sulfur source, a combination of sulfur sources has also been reported for the synthesis of MoS2 nanostructures. Luo et al.126 reported the synthesis of chiral MoS2 nanomaterials through a colloidal synthesis method. In this synthesis process, chiral cysteine enantiomers and NaHS both acted as the sulfur source, whereas MoO3 served as the molybdenum source. To ensure the formation of MoS2, ascorbic acid was introduced as a reducing agent. Moreover, the addition of cysteine enantiomers was responsible for the preferential folding of the MoS2 planes, leading to the formation of chiral MoS2 nanomaterials.
image file: d4qi03147d-f12.tif
Fig. 12 (a) Colloidal synthesis of MoS2 nanostructures using a one-pot heating method and (b) possible mechanism of nanostructure formation at different reaction temperatures. Reproduced from ref. 122 with permission. Copyright 2023, published by the American Chemical Society under a CC-BY license. (c–h) HRTEM images of MoS2 quantum dots capped with different concentration combinations of OAc and OAm. Reproduced from ref. 123 with permission. Copyright 2024, the American Chemical Society.

A series of Ag@MoS2 core–shell heterostructures was prepared via a two-stepped colloidal synthetic process (Fig. 13a).127 Initially, core@shell structured Ag2S@MoS2 heterostructures were formed via a one-pot procedure, in which the silver and molybdenum precursors were sequentially injected into the S powder containing oleylamine, respectively. The size of Ag2S@MoS2 can be adjusted by controlling the reaction time. Afterward, the core@shell Ag2S@MoS2 will be transferred to Ag@MoS2 via a facile heat treatment process in the presence of trioctylphosphine. Fig. 13b–e present the typical TEM images of the resulting samples. The Ag@MoS2 core–shell heterostructures demonstrate ideal platforms for real-time SERS application due to the strong electromagnetic field generated in the plasmonic Ag core. By simultaneously adding S and Se powder, Gong et al.128 prepared ultrathin MoS2(1−x)Se2x alloy nanoflakes. The chemical composition of the MoS2(1−x)Se2x alloy nanoflakes could be completely tuned. As presented in Fig. 13f–o, all the samples possessed a nanoflake morphology with monolayer or few-layer thickness and abundant edges. The dark-contrast lines represent the standing edges. An increase in the selenium content resulted in an increase in the lateral dimensions of the individual nanoflakes, growing from 10–15 nm (for x = 0 and 1/3) to nearly >20 nm (for x = 2/3 and 1). Furthermore, the nanoflake thickness also increased with an increase in the Se content, from mostly monolayers to predominantly few-layers. This result is ascribed to the stronger interlayer interaction at a high selenium content. Similarly, Li et al.129 presented a room temperature general colloidal method for the synthesis of TMD nanoparticles (including MoS2, RuS2, ReS2, MoSe2, RuSe2, and ReSe2). In their synthetic systems, the metallic chloride served as the metallic source and the combination of Na2S and NaHSe as the sulfide and selenide source, respectively.


image file: d4qi03147d-f13.tif
Fig. 13 (a) Synthesis of Ag2S@MoS2 core–shell heterostructure. (b) TEM and (c) HRTEM images of the synthesized Ag2S@MoS2 with a size of 25.8 ± 3.4 nm. (d) TEM and (e) HRTEM images of synthesized Ag@MoS2 with a size of 23.4 ± 3.8 nm. Reproduced from ref. 127 with permission. Copyright 2020, the American Chemical Society. TEM images of MoS2(1−x)Se2x nanoflakes with (f and g) x = 0, (h and i) x = 1/3, (j and k) x = 1/2, (l and m) x = 2/3, and (n and o) x = 1. Reproduced from ref. 128 with permission. Copyright 2015, the American Chemical Society.

The colloidal synthetic approach, overcoming the limitations associated with conventional strategies, presents a highly promising strategy for the large-scale production of MoS2, particularly nanoflakes. The main advantage of colloidal synthesis is the possibility of the tuning of the lateral sizes and stacking confinement of nanoflakes down to the atomic scale. However, its greatest disadvantage is the use of toxic precursors, such as CS2, TAA, and [Mo(CO)6]. Besides, these compounds are unstable in the presence of air or moisture in certain cases. Thus, in the future, less toxic compounds are necessary for this method. In addition, owing to the diversity of transition metals in crystalline phases, more attention should be paid to the phase control in colloidal synthesis in the future.130

2.4 Hydrothermal/solvothermal method

Hydrothermal synthesis is a conventional bottom-up approach, which is widely applied in synthetic chemistry. The hydrothermal method offers advantages of precious control of the crystal form to give a well-defined morphology. To date, a variety of functional materials has been successfully synthesized using this method, such as oxides, sulfides, nitride, and inorganic–organic hybrids. Geng et al.131 synthesized stable 2D metallic MoS2 (1T phase) nanosheets by mixing MoO3, thioacetamide (TAA) and urea in an autoclave and performing the reaction under hydrothermal conditions (200 °C). In this system, TAA serves as the S source given that its decomposition is capable of generating S2−, whereas urea plays the role of reducing agent to convert Mo(VI) into Mo(IV). The stacking of the layered structure in the resulting metallic MoS2 is dramatically decreased owing to the intercalation effect of water molecules. As a result, the resulting metallic MoS2 exhibits excellent stability in water. Kumar et al.132 prepared MoS2 micro/nanostructures, including microrods, microspheres, and microrods, using sodium diethyldithiocarbamate trihydrate as the S source. The resulting MoS2 microstructures possessed an increased interlayer spacing and excellent adsorption behavior for the removal of Pb(II) from aquatic systems. Saias et al.133 fabricated MoS2-supported alumina via a hydrothermal process and demonstrated the effect of the growth temperature on the crystallinity of MoS2 (Fig. 14a). At the screened temperatures, all the MoS2 samples possessed a flower-like morphology, which further assembled into ball-like microstructures (Fig. 14b–d). The growth temperature demonstrated a significant influence on the quality of the MoS2 flower spheres. Specially, a high temperature is favorable for forming spheres with a large diameter. This result is ascribed to the additional kinetic energy present during growth, which induces the formation of lower-energy thermodynamic products, exhibiting obvious boundaries formed on the top of the surface. To control the interlayer spacing in MoS2, Li et al.134 developed a hydrothermal process to prepare MoS2 submicrospheres assembled by tangled MoS2 nanosheets with the assistance of foreign species with varying sizes. As presented in Fig. 14e, three foreign molecules were examined, including L-cysteine, thiourea, and glucose. Fig. 14f–h display the corresponding HRTEM images of the samples, indicating their interlayer gaps are 0.62 (L-cysteine), 0.87 (thiourea), and 1.02 nm (glucose), respectively. Usually, foreign species with a larger size are favorable for producing a larger interlayer spacing. Obviously, glucose possesses the largest size among the three organic compounds, and thus the sample obtained with the introduction of glucose produces the largest interlayer gaps. L-cysteine is a small-sized molecule, producing interlayer gaps similar to that in pristine MoS2. Meanwhile, in this case, L-cysteine also serves as sulfur precursor to react with Na2MoO4, generating MoS2. In the case of thiourea, another small-sized molecule also acting as the sulfur precursor, it demonstrates a larger interlayer gap of 0.87 nm, which is increased by 0.25 nm compared to that of pristine MoS2. This result can be ascribed to the generation of NH3 molecules by the decomposition of thiourea, whose diameter is approximately 0.31 nm. Similarly, in the case of introducing glucose, the interlayer distance increased by 0.5 nm compared to that of pristine MoS2. This result demonstrates good agreement with the size of the monosaccharide, which is about 0.6–0.8 nm. In another work, Kaushik et al.135 utilized a similar process to prepare MoS2 with different morphologies, including nanoflower, interconnected nanoplates, and nanosheets (Fig. 14i–k), respectively. In their work, thiourea and ammonium molybdenum were used as the sulfur and molybdenum precursors, respectively, and morphology control was achieved by adjusting the reaction time, producing nanoflower (18 h in water), interconnected nanoplates (36 h in water), and nanosheets (48 h in MDF/water mixture). Yu et al.136 prepared rose-shaped MoS2 nanoflowers with TAA and Na2MoO4 as the precursors, demonstrating an efficient SER substrate for detecting R6G. Later, they further demonstrated that the rose-shaped MoS2 nanoflowers also are capable of serving as a support to deposit Au nanomaterials for the detection of crystal violet molecules.137 Panchu et al.138 applied the hydrothermal process to synthesize pristine MoS2 and an Ni-MoS2 composite. In their system, the formation of MoS2 is based on the reaction taking place between ammonium molybdate and thiourea. The interconnected sheet morphology of MoS2 is attributed to the unreacted NH4+ ions, which prevented the stacking of the nanosheets. Huang et al.139 reported a hydrothermal process, wherein ammonium molybdate, thiourea, and N-acetyl-L-cysteine (NAC) were used as the precursors and the capping reagent to obtain monolayer MoS2 quantum dots, which were water-soluble with a uniform lateral size of ≈2.1 nm.
image file: d4qi03147d-f14.tif
Fig. 14 (a) Procedures for the synthesis of MoS2 supported on an alumina substrate. (b–d) SEM images of MoS2 obtained at different temperatures. Reproduced from ref. 133 with permission. Copyright 2022, the American Chemical Society. (e) Synthesis of MoS2 microspheres with different interlayered spacings. (f–h) HRTEM images of MoS2 microspheres with different interlayered spacings. Reproduced from ref. 134 with permission. Copyright 2020, the American Chemical Society. SEM images of MoS2 with different morphologies: (i) nanoflower, (j) interconnected nanoplates, and (k) nanosheets. Reproduced from ref. 135 with permission. Copyright 2023, the American Chemical. Society.

Gao et al.140 developed a hydrothermal process to grow high-purity MoS2 square nanotubes. As presented in Fig. 15a, both Na2MoO4 and MnCl2 were employed in the system. During the initial stage, square-column MnMoO4 nanorods were formed via the reaction between Na2MoO4 and MnCl2. Owing to the presence of thiourea, the MnMoO4 nanorods could be converted into MnCO3 and MoS2, generating a core@shell-structured composite. Interestingly, the resulting MoS2 possessed a nanosheet morphology, covering the outside the composite. As the reaction proceeded, MnMoO4 gradually disappeared, generating the core@shell MoS2@MnCO3. Finally, the core MnCO3 was removed by washing with HCl, and thus MoS2 nanotubes were obtained. In this system, the formation of MnCO3 and MoS2 nanosheet can be ascribed to the decomposition of urea, generating CO32− and S2− under hydrothermal conditions. Fig. 13b–f show the SEM and TEM images of the resulting square MoS2 nanotube, which demonstrated the typical length of 7.13 μm and width of ∼634 nm. The interesting phenomenon in this system is the formation of MnCO3 but not MnS given that the former possesses a larger solubility product (Ksp (MnCO3) = 2.24 × 10−11) than that of MnS (Ksp (MnS) = 4.65 × 10−14).141,142 This phenomenon can be ascribed to the fact that the process is controlled by kinetics but not thermodynamics.


image file: d4qi03147d-f15.tif
Fig. 15 (a) Procedure for the formation of MoS2 nanotube. (b and c) SEM images of the resulting MoS2 nanotube at different magnifications. (d–f) Element mapping of square MoS2 nanotube. Reproduced from ref. 140 with permission. Copyright 2023, the American Chemical Society.

Similar to solid–gas synthesis, the hydrothermal process is also capable of yielding MoS2−xSex when both S and Se sources are employed in the system. For example, Taufik et al.143 prepared MoS2−xSex solid-solution nanoparticles using a hydrothermal process. The whole preparation process is presented in Fig. 16a. Initially, MoO42− is formed by dissolving Na2MoO4·2H2O in water, which then is reduced by hydrazine to generate Mo4+. Afterwards, the system is transferred to a hydrothermal vessel. Owing to the relatively high-temperature conditions, the Mo4+ ions will react with the sulfur and selenium powder, leading to the formation MoS2−xSex. The composition of the MoS2−xSex solid-solution nanoparticles can be preciously controlled by adjusting the selenium addition ratio. Fig. 16b–f display the TEM images of the MoS2−xSex samples. The size of these samples showed a decreasing trend with an increase in the selenium addition ratio. The as-x = 0 sample possessed a flower-like shape with a diameter of about 500 nm. The particle size of the as- x = 2 sample decreased to below 100 nm. In another work, Zhang et al.144 developed a two-step process to synthesize an MoS2−xSex/G composite alloy. As presented in Fig. 16g, MoS2 with an interlayer spacing of 0.62 nm was initially prepared via a hydrothermal procedure in the presence of graphene oxide. Afterward, a thermal selenization treatment is applied to convert MoS2/G to MoS2−xSex/G. Owing to the larger size of Se than S, MoS2−xSex produces a larger interlayer spacing (0.66 nm). The composition of MoS2−xSex in the final product can be conveniently controlled by adjusting the ratio of Se powder to MoS2/G composite.


image file: d4qi03147d-f16.tif
Fig. 16 (a) Hydrothermal process to synthesize MoS2−xSex. TEM images of (b) as-x = 0, (c) as-x = 0.2, (d) as-x = 1, (e) as-x = 1.8, and (f) as-x = 2. Reproduced from ref. 143 with permission. Copyright 2021, the American Chemical Society. (g) Schematic of two-step process to synthesize MoS2−xSex/G. Reproduced from ref. 144 with permission. Copyright 2019, the American Chemical Society.

Besides synthesizing pristine MoS2, composites containing MoS2 can also be obtained via the hydrothermal process. For example, Zheng et al.145 prepared an MoS2/CdS composite via a one-pot hydrothermal process. The resulting MoS2/CdS composite was rich in sulfur vacancies, which was verified by electron paramagnetic resonance (EPR) and UV–vis diffuse-reflectance spectroscopy (UV-DRS). When evaluated as a photocatalyst for the N2 reduction reaction, it delivered a high production of NH3 (249.7 mg L−1 g−1), which was 5.4- and 3.9-times higher than that of pure MoS2 (45.9 mg L−1 g−1) and pristine CdS (64.5 mg L−1 g−1), respectively. Similarly, Chen et al.146 prepared Co-doped MoS2 and CoS2 composites with a coupling interface via a one-pot hydrothermal process, which enabled the chemoselective formation of p-chloroaniline via the hydrogenation of p-chloronitrobenzene. By reacting pre-synthesized few-layer MoS2 with the ZnxCd1−xS/ZnS(en)1/2 heterostructure under solvothermal conditions with ethylenediamine as the solvent, Dong et al.147 prepared MoS2/ZnCdS/ZnS dual heterostructures, delivering ZnS nanosheets decorated with ZnCdS nanorods and few-layer MoS2. This advantageous heterostructure endowed the resulting sample with an excellent photocatalytic performance, offering an H2 evolution rate of up to 79.3 mmol g−1 h−1 under visible light irradiation. Wang et al.148 developed a two-step hydrothermal process to prepare a CuS/MoS2 composite. As presented in Fig. 17a, the synthesis involved two hydrothermal steps. It was noted that the composition of CuS/MoS2 could be adjusted in the second hydrothermal stage by controlling the amount of CuS. Fig. 17b–d present the morphology characterization of the sample obtained by using 0.75 mmol CuS, whose diameter is around 2 μm. Furthermore, the distance separating adjacent lattice fringes is 0.62 nm, which is in accordance with the (002) crystal plane of 2H-MoS2. In addition, the other observed lattice spacing of 0.28 nm is ascribed to the (103) crystal plane of CuS. Hu et al.149 prepared an MoS2/PANI composite via a hydrothermal process. Specially, they initially prepared PANI via chemical oxidation polymerization, which was further utilized as the substrate to deposit MoS2. For the formation of MoS2, MoO3 and KSCN were employed as the Mo and S sources, respectively. The resulting MoS2/PANI could be further converted to an MoS2/C composite. Both the MoS2/PANI and MoS2/C composites exhibited an excellent performance in lithium ion batteries.


image file: d4qi03147d-f17.tif
Fig. 17 (a) Hydrothermal synthesis of CuS/MoS2 composite. (b and c) TEM images and (d) high-resolution transmission electron microscopy (HRTEM) images of CuS/MoS2 composite. Reproduced from ref. 148 with permission. Copyright 2024, the American Chemical Society.

The hydrothermal process, using hot compressed water to convert reactants to the desired products, is emerging technology that contributes to the sustainable production of nanomaterials. Similar to colloidal synthesis, the hydrothermal process is also classed as a liquid synthetic system. Compared to other conventional processes, the reactions in the hydrothermal process are carried out in a closed system, leading to several advantages such as easy operation, energy efficiency, cost saving, better nucleation control, and less pollutant emission. Besides, the hydrothermal technique is capable of accelerating interactions between solid and fluid species, yielding phase-pure and homogeneous materials. Compared to CVD or milling, the hydrothermal process usually involves slightly a longer reaction time.

2.5 Salt-assisted synthesis

In the past few years, the salt-assisted synthetic technique has attracted wide attention for preparing many functional materials. Salts may play several roles in synthetic systems, such as the reaction medium, template for generating target materials with special shapes/porosity, and precursor.150,151 Cevallos et al.152 presented the growth of single crystals of MoS2 through a liquid salt flux made from a low-melting mixture of NaCl and CsCl. The crystal characterization revealed that the resulting MoS2 crystals were comprised of a dominant 2H-MoS2 phase with a very small percentage (about 3%) of 3R intergrowths. Suleman et al.153 report that NaCl is capable of promoting the formation of MoS2 via CVD, which is ascribed to the confinement effect of NaCl. Furthermore, the morphology of the grown MoS2 can be well controlled by adjusting the amount of NaCl, growth temperature, and Mo/S precursor ratio. Utilizing a similar NaCl-assisted CVD process, Singh et al.154 unlocked the role played by the reactant concentration in the boundary layer. Fig. 18a displays the diagram of the setup used for growing 3L-MoS2, whose critical parameters such as s, h, and d, are schematically shown in Fig. 18b. Here, s refers to the separation between the precursors and growing face of the SiO2/Si substrate, h refers to the height of the boat, and d is the distance between the growing face of the Si substrate and central axis of the tube/boundary layer. The gaseous precursors are input with the carrier gas (Ar) from a concentration boundary layer, as shown in Fig. 18c. To collect the formed MoS2, the SiO2/Si substrate was positioned on the crucible in a face-down manner. As a result, there was a space between the sidewalls of the crucible (Fig. 18d). The solid precursor of MoO3 was mixed with NaCl in the middle of the setup. Upon heating, the reactions could be confined in the space created by NaCl. The quality of the resulting MoS2, including layer number, growth area coverage, and nucleation density could be easily regulated by controlling the concentration boundary layer formation, which can be achieved by tuning the separation between the precursors and growing face of the substrate. Khan et al.155 demonstrated that the NaCl-assisted CVD process decreased the growth temperature from 850 °C to 650 °C and high-quality few-layer MoS2 could be prepared.
image file: d4qi03147d-f18.tif
Fig. 18 (a) Schematic of NaCl-assisted CVD process for 3L-MoS2 growth. (b) Schematic of definition of s, h, and d. (c) Reactant concentration variation in the synthesis system and its boundary layer formation in the CVD tube. (d) Schematic of A, B, and C points on the substrate. Reproduced from ref. 154 with permission. Copyright 2021, the American Chemical Society.

To reveal the mechanism of yielding MoS2 monolayers in the NaCl-assisted CVD system, Lei et al.156 employ theoretical tools, i.e., first-principles calculations and ab initio molecular dynamics (AIMD) simulations, to examine the growth of MoS2. According to the binary phase diagram of NaCl-MoO3 (Fig. 19a), there is a eutectic point at x(NaCl) = 0.59 and the system delivers a low melting point of 648 °C. This temperature is ∼150 °C lower than that of the pure compounds. Interestingly, the system can further be reduced to form MoO2Cl2 (dashed line in Fig. 19a). The formation of MoO2Cl2 can be described by the following equation:

 
3MoO3 (s) + 2NaCl (s) → MoO2Cl2 (g) + Na2Mo2O7 (l) (3)


image file: d4qi03147d-f19.tif
Fig. 19 (a) MoO3-NaCl binary phase diagram as well as chemical structure of MoO2Cl2. (b) Three plausible reaction pathways for MoO2Cl2 sulfurization toward MoS6. (c) Energy profiles of MoS2 monolayer growing process: salt-assisted (red line) and conventional (black line) CVD. Reproduced from ref. 156 with permission. Copyright 2022, the American Chemical Society.

Owing to its volatile feature, MoO2Cl2 reaches a significant vapor pressure (∼800 mbar) sufficient for experimental MoS2 growth at a temperature as low as 150 °C. In the presence of S2 gas (dominant sulfur species at the growth temperature), MoO2Cl2 can be sulfurized into MoSn (n = 4, 5, and 6) molecules, with MoS6 being the dominant one. Fig. 19b depicts the three plausible pathways. A comparison of the energy profiles in the salt-assisted synthesis and the conventional CVD process is provided in Fig. 19c. The former is ∼2 to 3 orders of magnitude faster at the growth temperature. Besides chloride, other halide salts such as F, Br, and I also deliver similar behavior in assisting the CVD growth of MoS2.

Besides acting as the reaction medium, some salts may serve as precursors to generate the target materials, which is dependent on the chemical properties of the utilized salts. Li et al.157 reported a KSCN molten salt strategy to synthesize MoS2 and MoSxSe2−x at a modest temperature of 320 °C. As presented in Fig. 20a and b, for both cases, the decomposition of KSCN is crucial, in which S2− is generated. For the formation of MoS2, an intermediate, MoS42−, is formed by reaction between MoO42− and S2−. Afterward, the formed MoS42− intermediate is pyrolyzed into MoS2 via heat treatment (Fig. 20a). For the formation of MoSxSe2−x, the Se powder may react with S2− released by the decomposition of KSCN to form the SexS2− polyanion, which further reacts with MoO42− upon heating, yielding MoSxSe2−x (Fig. 20b). Interestingly, the decomposition of KSCN is verified to be promoted by MoO42− to S2−, which then activates the Se powder, culminating in the formation of MoSxSe2−x. Furthermore, structural characterization revealed that the resulting MoSxSe2−x is characterized by abundant anion vacancies, which can be ascribed to the introduction of Se heteroatoms, causing lattice distortion. Compared to S2−, the large size of SexS2− offers substantial steric hindrance, resulting in the slow/uncompleted growth of the MoSxSe2−x crystal. Theoretical analyses indicated that the electronic structure of MoSxSe2−x can be well adjusted by Se atoms and anion vacancies. As a result, the optimized MoS1.5Se0.5 possesses a minimized band gap of 0.88 eV and an almost zero ΔGH* of 0.09 eV. Fig. 20c and d depict the IFFT patterns of MoS2 and MoS1.5Se0.5. The results indicated that MoS2 delivers well-defined lattice fringes, indicating high crystallinity (Fig. 20c). In the case of MoS1.5Se0.5, mainly an amorphous region was produced, exhibiting low crystallinity (Fig. 20d). Considering the high toxicity of KSCN, Jin et al.158 introduced Na2SO4 as the sulfur precursor to prepare MoS2. In their work, Na2SO4 also plays roles in adjusting the diffusion of the source precursors and balancing their mass flux. However, the system needs H2 to generate H2S, which reacts with the molybdenum source. Furthermore, the generation rate of H2S was demonstrated instantaneously. Meanwhile, the decomposition temperature of Na2SO4 is much higher (above 690 °C) than that of KSCN (below 350 °C).


image file: d4qi03147d-f20.tif
Fig. 20 Schematic of molten KSCN0assisted synthesis of MoS2 (a) and MoSxSe2−x (b). IFFT patterns of (c) MoS2 and (d) MoS1.5Se0.5. Reproduced from ref. 157 with permission. Copyright 2024, the American Chemical Society.

Compared to solid-state synthetic processes, salt-assisted synthesis shows advantages in terms of operation, demonstrating high reactivity for the preparation of various inorganic nanomaterials. In the case of molten salt systems with relatively low melting points, the ionic pool results in the high dispersion of the reactants, ensuring fast mass transfer. As a result, chemical reactions can be completed swiftly, showing high efficiency in generating products. In addition, molten salt systems have the advantages of low cost, high yield, time efficiency, and environmental friendliness. Furthermore, some particular salts were recently demonstrated to act as templates in the fabrication of inorganic nanomaterials with special morphologies and sizes, which have been applied to replace some of the traditional structure-directing agents or pore forming agents. These characteristics endow salt synthesis systems with a safer synthesis process, easy movability, and the possibility of scalable production. It should be noted that to date, the salt-assisted synthetic techniques are still in their infancy, which are subject to further investigation by emphasizing their general mechanisms in the future.

In this section, we discussed several common strategies. A comparison of these established techniques is presented in Table 1. It should be noted that many other methods are available to synthesize MoS2, such as atomic layer deposition,159–161 electron beam irradiation,162 and phase transition.163,164

Table 1 Advantages and disadvantages of methods for synthesizing MoS2
Method Advantages Disadvantages
Exfoliation Low pollution Complex technology
Low energy consumption Limited reproducibility
Wide applicability Large structural polydispersity
Chemical vapor deposition Large treatment capacity High equipment input
Wide applicability Complex technology
Flexible scale High energy consumption
Potential secondary pollution
Colloidal synthesis Capable of large-scale production Toxic precursors involved
Mild reaction conditions Poor crystalline control
Generates polluted liquid
Solvothermal Versatile precursors available Generates polluted liquid
Mild reaction conditions Relatively long reaction time
Energy saving  
Salt assisted synthesis High efficiency of electric heat conversion The technology is not mature
Low pollutant emissions  
Recovery of reusable materials  


3 CDI application

The search for high-performance electrodes constitutes one of the critical issues to realize CDI. To date, various materials have been demonstrated to be efficient CDI electrode candidates, such as carbon, MXenes, and PBA. Table 2 lists the advantages and disadvantages of these materials. Benefiting from their huge specific adsorption area, good conductivity and good chemical and thermal stability, carbon-based materials have been confirmed to be excellent CDI electrode candidates, such as activated carbon, carbon fibers, carbon aerogel, carbon cloth and graphene. However, a disadvantage of these carbon-based materials is their electric double-layer capacitance (EDLC) mechanism, delivering a low desalination capacity and poor selectivity. Metallic oxides mainly exhibit pseudocapacitive behavior, showing superior desalination capacity over carbon. However, the stability of metallic oxides constitutes the main barrier to their widespread application, which is very similar in capacitive energy storage systems. LDH, MXenes, and MoS2 belong to the class of 2D layered materials, which all suffer from easy restacking. Generally, the phase diversity of MoS2 offers more opportunities to adjust its phase crystal structure and form hybrids with other species to boost their CDI performance. In this section, we discuss the application of MoS2-based materials in CDI in detail.
Table 2 Advantages and disadvantages of some commonly available CDI electrodes
Electrode candidate Advantages Disadvantages
Carbon High electronic conductivity Limited by EDLC mechanism
Good stability Not all surfaces are fully utilized
High specific surface area  
Metallic oxides Variable oxidation state Poor cyclic stability
Pseudocapacitive mechanism  
High desalination capacity  
LDH Abundant active sites Poor electronic conductivity
High structural stability Ease of restacking
High desalination capacity Low desalination rate
MXene Excellent metallic conductivity Interlayer restacking
High hydrophilicity Easily oxidized
Ease of functionalization Hazardous agents involved in synthesis
MoS2 Adjustable phase to tune the conductivity Poor cyclic stability
Easy modification Ease of restacking


3.1 Pristine MoS2

Molybdenum disulfide has three different phase structures, and different phase types have different physical and chemical properties. The 1T phase is a special ultra-thin structure, with a large layer spacing, high conductivity, excellent electrochemical performance (>2H), high desalination performance, more active sites and higher ion migration rate than the 2H phase, but the thermal stability of the 1T phase is poor.165 Also, the 1T phase easily transforms into the 2H phase in the cycling process, and thus its synthesis process is complex. The 2H phase is a semiconductor with poor conductivity and strong van der Waals force between layers, resulting in its easy restacking. However, the 2H phase has high thermal stability and is suitable for long-term recycling. The 3R phase is a three-way stacked structure with small layer spacing, which is not conducive to ion diffusion, and its electrochemical performance has not been widely explored.166
3.1.1 Intrinsic MoS2. To date, both 1T and 2H phase MoS2 have been investigated for CDI application. In the case of the 1T phase, it usually demonstrates higher electronic conductivity (10–100 S cm−1) with a lower charge transfer resistance in electrochemical applications, leading to a superior desalination capacity and rate. This advantageous performance can be ascribed to the distorted tetragonal structure with band inversion and spin–orbit interaction. The greatest disadvantage of 1T MoS2 is its instability, tending to restack into the 2H phase and causing poor cycling stability. By contrast, 2H phase MoS2 is the most common and stable crystal structure, demonstrating semiconductor characteristics. When evaluated as an electrode in CDI cells, it delivered poor charge transport, suffering from rapid capacity decay and poor rate performance. Xing et al.14 reported that chemically exfoliated MoS2 (ce-MoS2) nanosheets demonstrate a comparable CDI performance to that of carbon materials and other common materials. The ion quality removal capacity reached 8.81 mg g−1, and the ion volume removal capacity reached 16.51 mg cm−3 for ce-MoS2. They further ascribed this excellent CDI performance of ce-MoS2 to the particular 2D layered thin sheet structure of the 1T phase, exhibiting excellent electrical conductivity and offering abundant space to accommodate ions. Meanwhile, the large interlayer spacing has also been demonstrated to be beneficial for the intercalation and transportation of ions. Wang et al.167 developed a flexible MoS2-based electrode with no binders by depositing MoS2 onto the framework of stainless-steel mesh (Fig. 21a). It was found that the surface treatment is beneficial for providing active nucleation sites for the growth of MoS2. The morphology observation revealed that MoS2 possessed Auricularia-like architectures. When evaluated as electrodes in a CDI cell, the binder-free MoS2-based electrodes delivered a better performance than that of pristine MoS2. The desalination capacity of the binder-free MoS2-based electrode was high as 28.76 mg g−1 (Fig. 21b and c). Considering that MoS2 exists in either the 1T or 2H phase, Ying et al.168 prepared a 1T-MoS2 electrode with a compact architecture via the restacking of exfoliated nanosheets, which exhibited a superior electrochemical performance to 2H-MoS2. The resulting 1T-MoS2 electrode demonstrated a high capacitance of up to ≈277.5 F cm−3 at an exceptional scan rate of 1000 mV s−1. For desalination application, the 1T-MoS2 electrode demonstrated an ultrahigh volumetric desalination capacity of 65.1 mg NaCl cm−3 in CDI experiments. Pang et al.169 prepared 1T-phase MoS2 with an expanded interlayer spacing of 9.8 Å, delivering an elevated capacitive contribution percentage (81%). When assembled into a CDI device, a high desalination capacity of 47.1 mg g−1 was achieved in a 200 mg L−1 NaCl solution with a fast desalination rate of 2.4 mg g−1 min−1. Zhang et al.15 investigated the removal of Pb(II) with a low concentration using an MoS2 electrode. They found that the removal mechanism involves two steps, electrosorption and chemisorption (CEC). Experimentally, they achieved an ultrahigh adsorption capacity at a voltage of 1.2 V, reaching 86.26 mg m−2. The mechanism investigation indicated that the enhanced adsorption capacity is mainly ascribed to two factors (Fig. 21d). On the one hand, the positively charged Pb(II) may undergo physical adsorption, which is achieved by van der Waals force given that Pb(II) can be attached and concentrated on the surface of molybdenite, which is achieved by EDL interaction through electrosorption. On the other hand, the exposed S atoms of molybdenite are capable of binding with Pb(II), leading to chemical adsorption. A high Pb(II)concentration in EDL is expected to promote the chemisorption of Pb(II) on the molybdenite electrode. Benefiting from this unique adsorption mechanism, the MoS2 electrode delivered a superior performance over graphite paper adsorbent, as evidenced in Fig. 21e. The maximum adsorption capacity reached 86.26 mg m−2. This value is nearly five-times that of single chemisorption without applying a voltage. Excitingly, the resulting molybdenite electrode also demonstrated potential for the CEC removal of other heavy metals. Taking Cd(II) as an example, CEC delivered a higher adsorption capacity (12.84 mg m−2) than chemisorption (3.87 mg m−2) (Fig. 21f).
image file: d4qi03147d-f21.tif
Fig. 21 (a) Procedure for synthesizing MoS2@SSM. (b) Schematic of the MoS2@SSM-based CDI cell. (c) CDI performance of MoS2@SSMB and MoS2@SSM. Reproduced from ref. 167 with permission. Copyright 2024, Elsevier. (d) Schematic for electroadsorption of Pb(II) on molybdenite electrode. (e) Comparison of isotherms of Pb(II) adsorption on graphite paper and molybdenite electrode. (f) Isotherms of CEC (with application of a voltage) and chemosorption (without applying a voltage) of Cd(II) on molybdenite electrode. Reproduced from ref. 15 with permission. Copyright 2018, Elsevier.
3.1.2 Elemental-doped MoS2. Heteroatom doping is an effective strategy to adjust the electronic structure of semiconductor nanomaterials, leading to improved optical or electrochemical performances.170–172 Li et al.173 developed a hydrothermal process to synthesize manganese-doped MoS2 nanosheets (Fig. 22a), which demonstrated an excellent CDI performance. Interestingly, manganese doping was demonstrated to enlarge the interlamellar spacing of MoS2, enriching its mesoporous structure. In the designed hybrid CDI cell, the salt removal capacity reached 24.5 mg g−1 in a 500 mg L−1 NaCl solution. Notably, the desalination process using Mn-MoS2 exhibited the characteristic of fast kinetics. The salt removal could reach the equilibrium state within min, and the removal rate was determined to be 3.06 mg g−1 min−1. Furthermore, no efficiency decay was observed after 30 cycles, demonstrating superior stability (Fig. 22b). Besides metallic cation doping, the incorporation of oxygen has also been reported as an effective approach to adjust the properties of MoS2. Sun et al.174 investigated the effects of incorporating oxygen in MoS2 nanosheets and concluded that it could dramatically boost the CDI performance of the MoS2 nanosheets. As presented in Fig. 22c, the optimized MoS2 electrode was determined to contain a moderate degree of oxygen (2.91 at%), exhibiting the highest desalination capacity of 28.85 mg g−1, which is almost 5.5-times that of MoS2 without oxygen. The mechanism investigation revealed that this enhancement can be ascribed to two factors. On one hand, owing to the high electronegativity of the incorporated oxygen atoms, the intrinsic conductivity can be dramatically improved. Simultaneously, more active sites will be exposed. Consequently, a higher specific capacitance and lower inner resistance may be generated when used as an electrode. On the other hand, the incorporated oxygen provides chances to achieve a balance between structure and active sites, endowing MoS2 with the highest salt adsorption capacity. The same group175 further demonstrated that O-doped MoS2 is capable of selectively recovering rare earth metals from a low concentration solution with a nearly 100% recovery efficiency. This high recovery efficiency is ascribed to the enhanced chemisorption and electrosorption process.
image file: d4qi03147d-f22.tif
Fig. 22 (a) Synthesis of Mn-MoS2. (b) CDI performance of Mn-MoS2. Reproduced from ref. 173 with permission. Copyright 2024, Elsevier. (c) Oxygen incorporation process and its effect on desalination capacity. Reproduced from ref. 174 with permission. Copyright 2020, Elsevier.
3.1.3 Defect engineering. The introduction of defects into nanomaterials demonstrates another versatile tool to boost their optical and electrical performances. Jia et al.176 developed defect-rich MoS2 sheets through thermal treatment. Owing to the abundant defects, MoS2 demonstrated an excellent desalination capacity. In particular, the desalination capacity reached 35 mg g−1, which is nearly three-times that of MoS2 without defects (12.8 mg g−1). This enhancement can be ascribed to several factors. Firstly, the defects on the surface create abundant negative charges, high specific capacitance and low inner resistance, which promote the electrostatic attraction and electrosorption of Na+ on the electrode surfaces. Secondly, defects can be incorporated on the surface of MoS2, which are beneficial for a superb CDI electrode performance. In another work, Zhan et al.177 reported the use of defective MoS2 as an electrode to capture Pd(II) through the CEC process. Fig. 23a–m depict the atomic configuration of defective MoS2 with variable vacancies (one or two). By combining theoretical and experimental investigation, they systematically examined the distinct adsorption behaviors of Pb(II) on the defective MoS2 electrode. Experimentally, a wide range of Pb(II) concentrations was examined (1–100 mg L−1). For sole chemosorption, an increasing trend was obtained in adsorption capacity versus the initial Pb(II) concentration. For example, it increased from 64.99 (at 8.37 mg L−1) to 159.98 mg g−1 (at 18.71 mg L−1). By applying a voltage, both the adsorption and removal efficiency showed a dramatically increase. Compared to the case of no applied voltage, the adsorption capacity and removal efficiency were boosted by almost 8 times, which can be attributed to the strengthened migration of ions, electrostatic interaction and electric double layer. Besides, for recycling use, the desorption of Pb(II) was also investigated, which could be completed within 60 min. After 10 cycles, the defective MoS2 electrode still maintained a removal of higher than 95%.
image file: d4qi03147d-f23.tif
Fig. 23 (a–m) Atomic configurations of defective MoS2 with one or two vacancies. Removal kinetics of Pb(II) on defective MoS2: (n) adsorption capacity and (o) removal. Reproduced from ref. 177 with permission. Copyright 2024, Elsevier.

The superior CDI performance of MoS2 can be ascribed to its excellent capacitive property, which originates from two aspects. Firstly, the good EDL capacitance delivered by MoS2 is owing to its sandwich-type S–Mo–S layered structure, which is capable of offering a high specific surface area. Simultaneously, the layered structure makes the intercalation of electrolyte ions in MoS2 easy, which is also beneficial for delivering an excellent capacitive property. Secondly, the oxidation state of the Mo atom can dramatically vary from +2 to +6, resulting in a good pseudocapacitive property. Elemental doping and introduction of defects can further vary the surface and electrochemical properties, leading to an enhancement in CDI performance. In particular, defect-rich MoS2 is capable of delivering abundant negative charges, high specific capacitance and low charge-transfer resistance, promoting the formation of a strong EDL in the CDI process and the electrosorption of Na+ on the surface of MoS2. Unfortunately, the restacking of the MoS2 sheets as well as the poor electrical conductivity of MoS2 hinder its application as a CDI electrode material. Thus, by compositing it with other species, the above-mentioned issues may be partly addressed. In the following sections, we discuss various MoS2-based composites for CDI application in detail.

3.2 Composites with carbon

In daily life, carbon materials are ubiquitous, which are usually characterized by rich porous structures and high specific surface areas. Currently, the application of carbon materials covers numerous fields, such as detection,178–181 adsorption,182–184 catalysis,185,186 energy and the environment.187–191 Similar to energy storage/conversion systems, the high conductivity, high surface area per unit mass, and intrinsic porosity make nanostructured carbon materials suitable candidates as CDI electrodes, such as activated carbon, carbon nanotubes, graphene, and carbon nanofibers. Also, the combination of carbon with inorganic pseudocapacitive candidates can dramatically boost its specific capacitance, and consequently the cycling stability of its inorganic counterpart will be improved. To date, numerous excellent works have been reported on the design of MoS2/carbon composites for CDI application. In this section, we provide a comprehensive summary on this sub-topic.
3.2.1 Activated carbon. Generally, activated carbon exhibits a particularly narrow pore size distribution, which has proven to be beneficial for increasing its salt adsorption capacity. However, the high internal resistance and physical pore blockage during extensive cycling result in a decay in its specific capacitance. Consequently, research has been dedicated to the incorporation of inorganic materials on the surface of carbon materials to enhance its specific capacitance through the combination of fast faradaic reactions and double-layer charging. Yin et al.192 reported the use of a carbon skeleton/molybdenum disulfide composite material (MoS2/AC) for the capacitive removal of Cd2+, achieving a removal capacity of 22.15 mg g−1. Utilizing the high affinity of PDA to inorganic nanomaterials, Hao et al.193 deposited few-layer MoS2 (FL-MoS2) on PDA spheres, which were prepared via the self-polymerization of DA under alkalic conditions. The obtained FL-MoS2/PDA spheres could be further converted into FL-MoS2/NCS (N-doped carbon spheres), as presented in Fig. 24a. In the composite, MoS2 possessed a hexagonal crystal structure. The presence of NCS prevented the growth of MoS2 along the (002) crystal plane, leading to a slight expansion of the interlayer distance. Besides, the introduction of NCS also improved the water wettability of the composite, as demonstrated by the lower water contact angle of FL-MoS2/NCS (62°) than that of the MoS2 electrode (96°). Benefiting from these features, the resulting FL-MoS2/NCS exhibited an excellent CDI performance for the removal of ions from water. Their results revealed that the FL-MoS2/NCS electrode possessed a high saturated Cu2+ adsorption capacity (1199.63 mg g−1) even at a low concentration (23.63 mg L−1). In particular, the competitive experiment (Cu2+/Na+/Fe3+) investigation revealed its good selectivity toward capturing Cu2+. In the case of the ternary system, the adsorption capacity for Cu2+ was 1071.6 mg g−1, which is much higher than that of Na+ (199.2 mg g−1) and Fe3+ (262.8 mg g−1). Several factors are responsible for the high selective adsorption of Cu2+. Firstly, the electrode/electrolyte wetting performance is improved by the excellent hydrophilicity of FL-MoS2/NCS, resulting in an improvement in its electrosorption performance. Secondly, the migration of ions can be dramatically accelerated owing to the enlarged interlayer spacing, abundant defects and mesoporous structure, ensuring fast adsorption kinetics. Furthermore, according to the hard soft acid–base theory, the adsorption of Cu2+ can be facilitated by the abundant S, N active sites (soft base) matching Cu2+ (soft acid). Applying XPS and competitive adsorption tests, they further revealed the Cu2+ adsorption mechanism. As presented in Fig. 24b, the FL-MoS2/NCS electrode provides two pathways to selectively capture Cu2+, electrical double layer (EDL) and complexation of Cu2+ with FL-MoS2/NCS. Besides, the Cu2+ adsorption data conforms to the pseudo-second-order kinetic model (Fig. 24c). Fig. 24d depicts the reusability results for the FL-MoS2/NCS electrode. A high capacity retention (94.5%) was retained after 5 cycles. Wu et al.194 reported the preparation of a hollow bowl-type carbon material loaded with MoS2 (HBC-MoS2), which also showed good selectivity for the removal of Cu2+ using the CDI technique. Similarly, its excellent performance is also ascribed to the synergistic effect of EDL and complexation between MoS2 and Cu2+. Nguyen et al.195 prepared N-doped carbon spheres using dopamine as the carbon precursor, which could be further utilized as a support to load flower-like MoS2 nanosheets. The obtained MoS2@NCS possessed a large specific surface area of 82 m2 g−1 and interconnected meso-macroporous channels, which were beneficial for rapid ion and electron transfer. The electrochemical investigation indicated that MoS2@NCS delivered a specific capacitance of 340 F g−1, demonstrating great potential in electrosorption applications. Further CDI measurement revealed that under the optimal conditions, MoS2@NCS delivered a superior specific electrosorption capacity of 59.9 mg g−1, in which the experimental data fits well with the modified Donnan model. The durability test indicated that the MoS2@NCS electrode showed no significant decline in desalination percentage even after 100 electrosorption-desorption cycles. Polyaniline (PANI) has similar properties to PDA, which can also be converted into N-doped carbon. Consequently, Liu et al.196 integrated MoS2 with carbonized PANI (MoS2/CP). Surprisingly, the interlayer spacing of MoS2 was dramatically enlarged due to the disordered entanglement between MoS2 and CP nanosheets, resulting in improved pore structure and surface area. These features are capable of offering multiple charge transfer routes and endowing more embedding sites for storing Na+. When evaluated as the electrode in a CDI cell, a remarkable desalination capacity (29.14 mg g−1) was delivered with a rapid desalination rate (2.9 mg g−1 min−1) and favorable cyclic durability. The mechanism investigation further revealed that the capacitance-controlled contribution accounted for 85.8%, whereas the diffusion-controlled contribution was 14.2%.
image file: d4qi03147d-f24.tif
Fig. 24 (a) Synthesis of FL-MoS2/NCS. (b) Schematic for the electrosorption of ions over FL-MoS2/NCS. (c) Time-dependent ASC and (d) cycling performance of FL-MoS2/NCS. Reproduced from ref. 193 with permission. Copyright 2020, Elsevier.

In another work, Tian et al.197 reported the preparation of hollow nano-flowered MoS2/N-doped hollow carbon spheres (MoS2/NHCS), which showed high selectivity for the removal of Pb2+ from wastewater using the CDI technique. As presented in Fig. 25a, the process for the preparation of MoS2/NCHS involves several steps. Initially, a core@shell SiO2@PDA composite is prepared via one-pot self-assembly, accompanied by polymerization and hydrolysis. This integrated process was also reported elsewhere.198,199 Afterward, PDA is carbonized into NC, followed removal of the SiO2 core via chemical etching. Consequently, NHCS is obtained, which can be further utilized as a support to deposit MoS2. Fig. 25b and c present the TEM and HRTEM images of the resulting MoS2/NHCS composite, which possessed a hollow structure and spherical morphology. Its hollow spherical edges are coated with lamellar structures. In addition, several lattice distances of 0.265, 0.226, and 0.153 nm were observed, which are attributed to the (101), (103) and (008) crystalline plane of the MoS2 crystal, respectively. When evaluated as a CDI electrode for the elimination of Pb2+, the resulting MoS2/NHCS demonstrated an excellent removal efficiency of 96.9% under the optimal conditions. Excitingly, excellent regeneration ability was achieved in 10 ppm Pb2+ solution, delivering 97.7% after 20 cycles, particularly in mixed solution to selectively remove Pb2+. This efficient selective elimination of Pb2+ can be ascribed to at least two aspects. Firstly, the formation of PbS precipitation promotes the preference of removing Pb2+ owing to the low solubility product of PbS, which was verified by in situ Raman spectroscopy and DFT calculations. Secondly, upon receiving electrons, some S ions can be transformed into S22−, leading to the generation of PbS2. Resorcinol formaldehyde resin can also be formed simultaneously with the hydrolysis of TEOS to generate SiO2@RF, which can be further treated similar to SiO2@PDA to generate hollow carbonaceous materials.200,201 As a result, Zhang et al.202 fabricated 1T-MoS2/C hybrid microspheres. When evaluated as a CDI electrode, the resulting 1T-MoS2/C hybrid microspheres delivered an excellent performance. The desalination capacity reached up to 48.1 mg g−1 at 1.2 V. In terms of kinetics, an outstanding desalination rate of 9.3 mg g−1 min−1 was achieved, which is much higher than that of the 2H-MoS2/C hybrid electrode (3.2 mg g−1 min−1). Besides, the energy consumption evaluation indicated not only a lower energy consumption (0.61 kW h kg−1) but also the higher charge efficiency (0.91) of the 1T-MoS2/C hybrid electrode. Furthermore, an exceptional stability performance was also obtained, preserving a high salt adsorption capacity of 41.5 mg g−1 after 20 cycles. This high CDI performance can be ascribed to the superior structure of the hollow carbon microspheres, offering the confinement effect. On one hand, the presence of hollow carbon microspheres restricts the overgrowth and agglomeration of MoS2 nanosheets. In this way, the active sites of MoS2 can be fully exposed, leading to a high removal capacity and rate. On the other hand, the stability of the whole electrode material can be dramatically improved.


image file: d4qi03147d-f25.tif
Fig. 25 (a) Synthesis of MoS2/NHCS. (b) TEM image and (c) HRTEM images of MoS2/NHCS. (d) Pb2+ removal efficiency of MoS2/NHCS at different concentrations of Pb(NO3)2. (e) Removal efficiency of heavy metal ions with MoS2/NHCS in a multicomponent solution. Reproduced from ref. 197 with permission. Copyright 2024, Elsevier. (f) Procedure for synthesizing MoS2/NMOC composite. (g) Water contact angles of MoS2, NMOC, and MoS2/NMOC. (h) CDI performances of MoS2, NMOC, and MoS2/NMOC. Reproduced from ref. 203 with permission. Copyright 2021, Elsevier.

In another work, Tian et al.203 utilized a similar principle to prepare an MoS2/NOMC (nitrogen-doped highly ordered mesoporous carbon) composite, as schematically shown in Fig. 25f. Owing to the introduction of NMOC, the presence of oxygen-containing groups improved the water wettability, which was evidenced by the water contact angle (Fig. 25g). The contact angles of the pristine MoS2 and NMOC are 96.2° and 85.1°, respectively. Unexpectedly, MoS2/NMOC produced a contact angle of 65.1°. This dramatically enhanced hydrophilicity can be ascribed to the generation of unpaired electrons through nitrogen doping, which induced the formation of some defects and disorder on the MoS2/NOMC composite surface.204 The electronic conductivity investigation revealed that the introduction of NMOC also improved the conductivity of the MoS2/NMOC composite. Owing to the above-mentioned features, the diffusion of salt ions into the MoS2/NOMC electrode surface was dramatically improved, leading to an enhanced desalination performance. As presented in Fig. 25h, the electrosorption capacity of the MoS2/NOMC electrodes reached 28.82 mg g−1 at 1.6 V, which is superior to that of NMOC and pure MoS2. The mechanism investigation indicated that the desalination process is mainly controlled by the capacitor process (with the contribution of 73.17%), whereas the diffusion-controlled process only contributed 26.83%.

Considering the semiconductive feature of MoS2, it is feasible to couple CDI with photocatalysis to remove some organic pollutants from water. Li et al.205 designed a system for eliminating tetracycline (TC), in which MoS2/C plays dual functions, i.e., photocatalyst and electrode, for CDI (Fig. 26a). In their work, the MoS2/C composite was prepared via the one-pot hydrothermal treatment of fructose, Mo source and S source. Owing to the introduction of fructose-derived carbon, the conductivity of the composite was improved compared with that of pure MoS2, and the recombination of photo-generated electron–hole pairs will also be efficiently prevented, which is beneficial for the photocatalytic degradation of TC, as presented in Fig. 26b. Experimentally, the TC removal capacity reached up to 1302.5 mg g−1 and the removal efficiency was 84.8%. The kinetic investigation indicated that the system offers fast removal kinetics. Moreover, no activity loss was observed even after six cycles of photocatalytic degradation and electrosorption (Fig. 26c–e).


image file: d4qi03147d-f26.tif
Fig. 26 (a) Schematic of MoS2/C-based system coupling with CDI and photocatalysis. (b) Photocatalytic mechanism for the degradation of TC over MoS2. (c) Removal kinetics for TC in the coupled system. (d) Initial concentration-dependent removal capacity for TC over MoS2/C0.5 electrode. (e) Cyclic stability of MoS2/C0.5 electrode for the removal of TC after six cycles. Reproduced from ref. 205 with permission. Copyright 2024, Elsevier.
3.2.2 Carbon nanotubes. Carbon nanotubes (CNTs) possess the typical 1D structure with diameters of a few nanometers. Structurally, CNTs consist of helical microtubules of graphitic carbon. The sp2 hybridization endows CNTs with fascinating thermal and electrical properties, showing great promise in electrochemical applications.206–209 The intrinsic tube cavity in CNTs offers paths for ion transfer, demonstrating a superior average salt adsorption rate. When compositing with MoS2, the extended axial ratio and 1D morphology of CNTs increased the contact area with MoS2 sheets. As a result, the interfacial resistance between the two may be reduced, increasing the active sites for ion adsorption. Srimuk et al.210 fabricated MoS2/CNT binder-free electrodes, enabling a stable desalination performance over 25 cycles in various molar concentrations, with salt adsorption capacities of 10, 13, 18, and 25 mg g−1 in 5, 25, 100, and 500 mM NaCl aqueous solutions, respectively. By assembling β-CD and micelles (consisting of the nonionic surfactant P123 and sodium oleate) in a dispersion of CNTs, Cai et al.211 obtained a CNT-β-CD complex, in which a 3D architecture was constructed by the carbon nanotubes (CNT) and β-CD. The CNT-β-CD complex was further used as a substrate to load MoS2, forming the MoS2@CNT-β-CD composite. After carbonization treatment, the β-CD could be converted into carbon spheres (CS). Consequently, 2D MoS2 connected with 3D carbon spheres (CS) (MoS2@CNT-CS) was obtained, demonstrating an excellent CDI performance (Fig. 27a). Owing to the highly dispersive characteristics of MoS2@CNT-CS, it provided sufficient effective faradaic sites for ion storage and transfer, delivering an enhanced SAC (25.35 mg g−1), fast SAR (3.9 mg g−1 min−1), and satisfactory cycling stability. Particularly, the specific ion selectivity evaluation of the MoS2@CNT-CS electrode revealed a selectivity order of Ca2+ (0.43 mmol g−1) > Na+ (0.36 mmol g−1) > Mg2+ (0.28 mmol g−1) > K+ (0.19 mmol g−1), as evidenced in Fig. 27b. Unexpectedly, the MoS2@CNT-CS electrodes showed no structure change after the electrosorption of metal ions, as demonstrated by XRD (Fig. 27c). The molecular dynamic (MD) calculations further provided evidence for the selectivity order of the screened ions. As presented in Fig. 27d–g, the bond parameters indicate that the intercalation energy follows the order of Ca2+ (−2.55 eV) < Na (−1.56 eV) < Mg (−1.37 eV) < K (−1.00 eV).
image file: d4qi03147d-f27.tif
Fig. 27 (a) Synthesis of MoS2@CNT-CS and CDI profile. SAC (b) of CNT-CS//MoS2@CNT-CS cell. Postmortem XRD curves (c) of MoS2@CNT-CS electrode after the saturated electrosorption of different cations. Atomic configuration of Na atoms (d), K atoms (e), Ca atoms (f) and Mg atoms (g) between MoS2 interlayers. Reproduced from ref. 211 with permission. Copyright 2021, Elsevier.

Due to the high electrical conductivity and excellent mechanical strength of carbon nanotubes, when they are combined with MoS2, thy will form a three-dimensional conductive network, and at the same time, as a substrate to support MoS2, inhibit its self-aggregation. Consequently, it enhances the electrical conductivity and mechanical properties of the composite material, and then improves the electrochemical performance of the material. However, the synthesis of CNTs is complicated, which is attributed to the high mechanical strength of CNTs, making their dispersion in solution difficult, and thus their preparation is complex and costly.

3.2.3 Graphene. Graphene is a two-dimensional sheet material with a 1-atom thickness. Owing to its conjugated structure, graphene exhibits excellent ionic conductivity, and thus has found wide application in the CDI desalination process. However, the π–π interaction between graphene sheets can cause reaggregation, reducing the available surface for ion adsorption, which becomes the major obstacle for its use in the field of CDI. Recently, the assembly of graphene onto other species, such as metal oxides and sulfides, heterogeneous atom-doped carbon, and surface charged carbon nanostructures, has been demonstrated to be an optimal strategy to improve the performance of CDI electrodes. Meanwhile, incorporating graphene into MoS2 is capable of expanding the distance between the stacked adjacent MoS2 layers, improving the mass transport and electron transfer. Han et al.212 reported the fabrication of an MoS2–graphene hybrid electrode for CDI applications (Fig. 28a and b). In this MoS2–graphene hybrid, graphene was employed as the substrate to support numerous MoS2 pleats. This configuration is beneficial for increasing the specific surface area and exposing more active sites. As presented in the TEM (Fig. 28e–f) and high-resolution transmission electron microscopy (HRTEM) (Fig. 28g) images, the surface of the graphene was homogeneously covered by MoS2 nanosheets. Owing to the relatively separate states between graphene and the MoS2 nanosheets, the interlayer spacing in the MoS2 nanosheets showed no obvious expansion, which is 0.68 nm (Fig. 28g), in good agreement with the (002) lattice plane of the hexagonal MoS2. Benefiting from the above-mentioned characteristics together with conductive networks formed by graphene, the MoS2–graphene hybrid electrode delivered a high volumetric NaCl adsorption capacity (VACNaCl) and high gravimetric NaCl adsorption capacity (GACNaCl) as a result of the high electrical conductivity of graphene and the rapid ion transport of MoS2. In another work, Gao et al.213 reported the preparation of an MoS2/rGO composite, which can be used as intercalation electrode for CDI. Experimentally, a remarkable salt capacity of 34.20 mg g−1 was delivered with a high charge efficiency of up to 97% in 300 mg L−1 NaCl aqueous solution using MoS2/rGO as the electrode (Fig. 28h–k). They also confirmed the role of the incorporated rGO in the composite. On one hand, rGO serves as a conductive support, ensuring fast electron transfer. On the other hand, the structural characterization indicated that the interlayer spacing of MoS2 was expanded to 0.73 nm after incorporating rGO, which was only 0.62 nm without rGO. Owing to the widened MoS2 interlayer, the diffusion of cations could be dramatically accelerated, while the internal strain decreased. In addition, rGO created a large number of available active adsorption sites and space to accommodate cations from the electrolyte.
image file: d4qi03147d-f28.tif
Fig. 28 (a) Schematic of CDI device; (b) schematic of preparation of MoS2–graphene hybrids; (c and d) SEM images of MG-1.6; (e and f) TEM images of MG-1.6; and (g) HRTEM image of MG-1.6. Reproduced from ref. 212 with permission. Copyright 2019, the American Chemical Society. (h) Plots of SAC vs desalination time in 300 mg L−1 NaCl solution at 1.4 V with a flow rate of 12 mL min−1. (i) Corresponding Ragone Plots. (j) SAC at different voltages in 300 mg L−1 NaCl solution of the MoS2//AC and MoS2/rGO//AC HCDI systems. (k) SAC and charge efficiencies of MoS2/rGO//AC at a voltage of 1.2 V with different initial NaCl concentrations. Reproduced from ref. 213 with permission. Copyright 2020, the American Chemical Society.

Liu et al.214 developed an MoS2/graphene oxide heterojunction (MoS2/GO-H), demonstrating an effective electrode for CDI. As presented in Fig. 29a, a single hydrothermal process was utilized to synthesize the MoS2/GO-H composite. The resulting MoS2/GO-H played dual roles in the removal of UiO22+, i.e., electrosorption and electrocatalysis (Fig. 29b). The morphology observation indicated that the surface roughness of MoS2/GO-H significantly increased compared to the original materials given that the MoS2 layers were seamlessly woven into the GO sheets. Fig. 29c presents the HRTEM image of MoS2/GO-H, revealing the crystalline spacing of ≈0.62 nm, which corresponds to the (002) crystal plane of MoS2. Compared to pristine MoS2, the introduction of GO-H resulted in a large increase in specific surface area (from 10.64 to 16.52 m2 g−1) and pore volume (from 0.025 to 0.124 cm3 g−1), as evidenced by N2 adsorption–desorption measurement (Fig. 29d). When evaluated as a CDI electrode for removing UiO22+ from water, MoS2/GO-H delivered an excellent performance in terms of removal capacity, adsorption kinetics, selectivity and recycling durability (Fig. 29e–j). Under the optimal conditions, the adsorption capacity reached 805.57 mg g−1 within ∼240 min with over 90% removal efficiency in the presence of numerous competing ions. Notably, MoS2/GO-H delivered a high desorption rate of 74% after five CDI cycles. Furthermore, its adsorption capacity retention was about ≈70%, indicating its excellent recycling performance.


image file: d4qi03147d-f29.tif
Fig. 29 (a) Synthesis of MoS2/GO-H. (b) In situ electrolytic deposition with complexation strategy for UO22+. (c) HRTEM image and (d) N2 adsorption–desorption curve of MoS2/GO-H. (e–j) CDI performance of MoS2/GO-H in the removal of UO22+. Reproduced from ref. 214 with permission. Copyright 2024, Wiley-VCH.

Zheng et al.215 reported the synthesis of an rGO@PEI/MoS2 composite rich in heterostructures. The interaction between rGO and MoS2 to form the rich heterostructures was verified to be van der Waals forces. Owing to the rich heterostructures, the rGO@PEI/MoS2 composite delivered a maximum specific capacitance of 212.53 F g−1, which was nearly a 45% improvement compared to that of rGO/MoS2. When evaluated as an electrode in a CDI cell, owing to the synergistic effect generated by the rich heterostructure, the rGO@PEI/MoS2 electrode delivered a desalination capacity of about 24.13 mg g−1. Peng et al.216 prepared 3D flower-like MoS2/rGO composites using a water/ethanol mixture as the solvent via a one-step hydrothermal method. With a water/ethanol ratio of 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v), the resulting MoS2/rGO composite possessed a big BET surface area (29.7 m2 g−1), which is beneficial for its CDI performance. The desalination capacity in 200 mg L−1 NaCl solution was determined to be 16.82 mg g−1 at 1.0 V. Taking advantage of chemical etching, the same group217 further fabricated MoS2/reduced graphene oxide (rGO) composites in which MoS2 contained defects. As presented in Fig. 30a, chemical etching was achieved based on NaBH4 reducing MoS2. The MoS2 defect had a significant effect on the structure of the composite and its electrochemical and desalination performances. Taking advantage of the fewest amount of (002) interplanar layers, the defect-containing MoS2/rGO composite (dMSG) could be obtained by introducing NaBH4 as a reductant. Meanwhile, the number of defects was dependent on the amount of NaBH4. In their work, the optimal amount of NaBH4 was found to be 5.0 g, and the corresponding sample was marked as dMSG-5, which delivered the smallest water contact angle and zeta potential, the largest specific capacity (305.6 F g−1 at 1.0 A g−1), and the lowest interfacial charge transfer resistance. When used for the CDI cell, dMSG-5 showed the largest desalination capacity, which was determined to be 25.47 mg g−1 at the voltage of 0.8 V in 200 mg L−1 NaCl solution. The mechanism survey indicated that the greater number of adsorption sites for Na+ insertion/extraction and ion diffusion channels was responsible for the improved ion accessibility and transport efficiency, leading to an enhanced CDI performance. Besides, regeneration of the dMSG-5 electrode could be easily achieved, and its conductivity could rapidly recover, indicating the excellent performance of electrosorption-desorption in the processes of regeneration (Fig. 30b). Excitingly, the electrosorption capability of Na+ showed no significant loss compared to the first electrosorption (from 25.47 mg g−1 to 24.79 mg g−1) after the 5th cycle (Fig. 30c).


image file: d4qi03147d-f30.tif
Fig. 30 (a) Schematic for the chemical etching of MSG with various dosage of NaBH4 and the electrosorption of Na+ in the EDL of the dMSG composite electrodes. (b) Electrosorption-desorption cycle curves and (c) cyclic electrosorptive capacity of Na+ on dMSG-5 electrode. Reproduced from ref. 217 with permission. Copyright 2021, Elsevier.

The combination of graphene and MoS2 demonstrates several benefits in boosting the CDI performance of MoS2. On one hand, the conductivity will be improved, which is ascribed to the high conductivity of carbon materials. On the other hand, the surface functional groups of carbon materials will improve the hydrophilicity of MoS2, resulting in an enhancement in the ion migration kinetics. In addition, the flexible adjustment of the porous structure of carbon offers further opportunities to control the ion adsorption behavior. However, a disadvantage is its poor selective adsorption capacity, which may constitute a future research direction in this area.

3.2.4 Carbon nanofibers. In the past few decades, carbon nanofibers, as a type of amorphous carbon with high porosity, high specific surface area, and excellent electric and thermal conductivity, have been regarded as the most promising conducting 1D material. Furthermore, the exposed stacking arrangements of graphene sheets and edge plane defects allow efficient electron transfer on the outer wall of CNFs. The advantages of excellent structural stability and high electrical conductivity endow carbon nanofibers with an excellent performance in supercapacitors, capacitive deionization, and oil/water separation. They can be used as conductive scaffolds and as self-supporting electrode materials. By incorporating carbon nanofibers into inorganic materials, they can act as conductive scaffolds and as self-supporting electrode materials, creating electron transfer pathways. As a result, the electrochemical performance of intercalation materials can be improved to a great extent due to the synergistic interaction between two species. Liu et al.218 developed MoS2 nanoflake-coated carbon nanofibers (CNFs@MoS2) through electrospinning and subsequent hydrothermal reaction (Fig. 31a). By changing the ratio of CNFs in the composite, the interlayer spacing of the (002) and (100) crystal planes of MoS2 could be adjusted. As presented in Fig. 31b–d, the layer spacing for CNFs@MoS2-1, CNFs@MoS2-2, CNFs@MoS2-3 was measured to be 0.50/0.21, 0.51/0.24, and 0.48/0.22 nm, respectively. Owing to its dual-mode capacitive behavior, CNFs@MoS2 delivered excellent desalination performances with a desalination capacity of 53.03 mg g−1 and desalination rate of 0.157 mg g−1 s−1 as well as good long-term stability (11.2% reduction in 30 cycles) (Fig. 31e–g).
image file: d4qi03147d-f31.tif
Fig. 31 (a) Procedure for preparing CNFs@MoS2 composite. HRTEM images of CNFs@MoS2-1 (b), CNFs@MoS2-2 (c), and CNFs@MoS2-1 (d). (e) Desalination capacities/charge efficiencies of the CNFs@MoS2-2-based RCDI at various currents. (f) Desalination capacity of the CNFs@MoS2-2-based RCDI system at various initial concentrations. (g) Cycling performance and charge efficiency of the CNFs@MoS2-2-based RCDI system. Reproduced from ref. 218 with permission. Copyright 2021 Elsevier.

This section mainly introduces hybrid materials composed of molybdenum disulfide and four common carbon-based materials. It is understood that the properties of different carbon-based materials vary greatly after forming composites. AC has a high specific surface area, high porosity, and can provide abundant ion transport channels, with low cost and a wide range of raw materials. However, its poor conductivity and uneven pore size distribution result in a poor and unstable electrochemical performances by composite materials. CNTs have high electrical conductivity and mechanical strength, but they are prone to aggregation and have high synthesis costs. CNFs also have excellent conductivity, controllable structure, high mechanical strength, but low specific surface area and complex synthesis process. Graphene has ultra-high conductivity, and using GO as a substrate can prevent the suppression of MoS2 self-stacking, expand the interlayer spacing, and expose more active sites. However, it is prone to aggregation, complex to prepare, costly, and has weak interface bonding between the two, resulting in average material stability.

3.3 Composite with conducting polymer

Besides compositing with inorganics, conducting polymers have also been reported to boost the CDI performance of MoS2. For example, Wang et al.219 demonstrated that the incorporation of PDA may improve the water wettability of MoS2/PDA, leading to lower inner resistances, higher specific capacitance, enhanced electrosorption rate and desalination capacity. Chen et al.220 developed a core/shell structured PPy@MoS2 composite, exhibiting an excellent CDI performance. As presented in Fig. 32a, the synthesis of PPy@MoS2 started from the formation of MoO3 nanorods under hydrothermal conditions with the assistance of HNO3, followed by the oxidative in situ polymerization of Py to generate the MoO3@PPy composite. It was noted that PPy was coated on the surface of MoO3 in the MoO3@PPy composite. Finally, to convert MoO3@PPy into PPy@MoS2, a further sulfurization treatment was necessary, which was carried out under hydrothermal conditions using thiourea as the sulfur source. Interestingly, MoS2 was grown on the surface of PPy. This phenomenon was also observed in the thermal treatment of FeOOH@PPy to generate the Fe3O4@NC composite.221 The crystalline phase investigation revealed that MoS2 existed as a hybrid phase of both 1T and 2H phases. Fig. 32b and c present the SEM and TEM images of the resulting one-dimensional PPy@MoS2, respectively. Obviously, the MoO3 microrods were replaced by PPy microtubes, which further served as the substrate for growing MoS2 nanosheets, leading to MoS2 being tightly adhered on the PPy microtubes. Benefiting from these features, the resulting PPy@MoS2 is capable of offering more active sites, boosting the desalination performance by reducing the surface energy. Fig. 32d and e display the CDI performances of PPy, MoS2, and PPy@MoS2, from which it can be concluded that PPy@MoS2 delivered a superior performance to pristine PPy and MoS2. In particular, a high desalination capacity of 159.6 mg g−1 was achieved by PPy@MoS2 with a high rate of 3.2 mg g−1 min−1. This outstanding performance can be attributed to the core–shell structure and hybrid phases, which ensure excellent charge efficiency, cycling stability, and low energy consumption (0.22 kW h kg−1), demonstrating its potential for practical desalination applications.
image file: d4qi03147d-f32.tif
Fig. 32 (a) Procedures for the synthesis of PPy@MoS2. (b) SEM and (c) TEM images of PPy@MoS2. Comparison of the desalination capacity of PPy, MoS2, and PPy@MoS2 at different current densities (d) and different potentials (e). Reproduced from ref. 220 with permission. Copyright 2024, Elsevier.

Shehzad et al.222 developed orderly mesoporous composite materials consisting of carbon, PPy and MoS2. Owing to the introduction of porous activated biocarbon (BCL) and PPy, the resulting composite exhibited the improved diffusion of hexavalent uranium ions owing to the highly conducive network, promoting electron transfer during the CDI process. Under the optimal conditions, an inspiring sorption capacity of 417.90 mg g−1 was obtained at pH = 4.5 and applied voltage of −0.9 V, which is better than that of many other materials used in CDI applications having a similar geometry and morphology. N,P-co-doped templated biochar (BBC) was obtained via the pyrolysis of Bambusa vulgaris Culm, and a BBC/2D-MoS2/PPy hybrid CDI system was assembled, which delivered a sorption capacity of 402.13 mg g−1 for the selective removal of U(VI).223 The same group224 also reported the fabrication of an MoS2/CNT/PPy-based composite electrode for U(VI) electrosorption, obtaining an experimental sorption capacity of 274.50 mg g−1 at − 0.90 V. Fang et al.225 developed a ternary composite comprised of DBS ion-doped polypyrrole (PPy-DBS)-coated MoS2 nanosheets, which were tightly attached to hollow carbon spheres (HCS). In this composite, the dispersion of MoS2 nanosheets could be dramatically improved by HCS owing to the increased layer spacing. Consequently, abundant active sites were provided for ion intercalation and transfer. In addition, PPy-DBS may serve as an electron transport path, ensuring fast on migration kinetics and excellent electrical conductivity. Owing to these advantages, the specific capacitance of the resulting PPy-DBS-MoS2@HCS reached 168.77 F g−1, which is much higher than that of MoS2 (72.9 F g−1). Coupling with HCS, an asymmetric CDI cell, HCS//PPy-DBS-MoS2@HCS, was further constructed to evaluate the CDI performance of PPy-DBS-MoS2@HCS. An extremely high desalting capacity (53.4 mg g−1) was obtained at 1.2 V in 500 mg L−1 NaCl solution. This salt adsorption capacity is 2.9- and 2.4-times higher than that of HCS//MoS2 (18.3 mg g−1) and HCS//MoS2@HCS (22.3 mg g−1), respectively. Liu et al.226 developed a CNT/PPy/MoS2 ternary composite, achieving a maximum desalination capacity of 24.8 mg g−1, ultra-high deionization rate of 5.24 mg g−1 min−1, and superior capacity retention rate of 92.7% over 25 cycles. Zargar et al.227 reported the preparation of a Ti3C2Tx/1T-MoS2/CNT (TMC) heterostructure with an interconnected network, delivering an excellent CDI performance. As presented in Fig. 33a, they initially prepared Ti3C2Tx/CNTs via self-assembly with the assistance of ultrasonic irradiation and stirring. Owing to the negative charge of the MXene surface, the CNTs could be easily dispersed in the solution well. Afterward, a one-pot hydrothermal procedure was carried out to deposit MoS2 onto Ti3C2Tx/CNTs, forming the TMC heterostructure. The composition and structural characterization indicated that the Ti3C2Tx MXene possessed an accordion-like morphology and the MoS2 material possessed a nanoflower-like morphology with 1T phase. The introduction of CNTs allowed the fabrication of a well-defined 3D-interconnected network by connecting MoS2 petals and MXene isles. Owing to these characteristics, the TMC heterostructure was capable of constructing multiple routes for efficient electron transfer, elevating the overall electrical conductivity. When used as the electrode in a hybrid CDI cell coupled with activated carbon, a remarkable salt desalination performance was achieved, including a high salt removal capacity (25.82 mg g−1), outstanding high removal rate (5.25 mg g−1 min−1), and superior cyclic stability (98.1% desalination capacity retention rate). It was noted that these parameters are all superior to that of single MXene and MoS2 as well as Ti3C2Tx/MoS2 (TM)-based materials, as evidenced in Fig. 33b–e.


image file: d4qi03147d-f33.tif
Fig. 33 (a) Schematic of the fabrication of MXene/MoS2/CNT composites. (b) Conductivity and (c) desalination capacity versus time, (d) CDI Ragone plots of the prepared electrodes, (e) adsorption capacities of TMC composite at different voltages (0.8, 1, and 1.2 V) and different initial NaCl concentrations. Reproduced from ref. 227 with permission. Copyright 2024, Elsevier.

In summary, integrating conducting polymers into MoS2 nanosheets presents a promising solution for boosting their CDI performance, providing crucial benefits such as hydrophilicity, adjustable interlayer spacing, and abundant active sites. These characteristics enable the direct uptake of ionic pollutants in water and synergistic adsorption of ions from water. The enhancement mechanism of combining conducting polymers with MoS2 can be ascribed to several factors. Initially, the conducting polymer may act as a supportive framework to inhibit the agglomeration of MoS2 nanosheets, allowing full contact with salt solutions. This solution wettability improvement is also partly attributed to the abundant functional groups in conducting polymers. Secondly, the introduction of a conducting polymer may diminish the charge transfer resistance, which can be ascribed to the good electronic conductivity of conducting polymers. Furthermore, conducting polymers are expected to supply additional electrochemical active sites, boosting the electrosorption capacity. In addition, the interspacing of MoS2 can be enlarged owing to the presence of a conducting polymer, which is evidenced to be beneficial for improving its CDI performance in terms of mass/charge transfer kinetics.

3.4 Composites with oxides

Many metal oxides, including hydroxides, possess a large specific surface area, good chemical stability, and huge specific capacitance value, making them promising electrode candidates in CDI systems. To date, some reports have demonstrated the potential applications of metal oxides with good adsorption properties in the area of CDI. In particular, the combination of metal oxides and 2D layer compounds is a versatile approach to enhance the electrical conductivity of electrode materials, leading to improved electrochemical and CDI performances. Yao et al.228 reported a two-step hydrothermal process to fabricate MoS2/MnO2 composites (Fig. 34a), which leveraged the atomic-level “pump-driving” effect. Owing to the charge redistribution catalyzing IEF, the electron/ion transfer capabilities within the MoS2/MnO2 heterostructure were dramatically enhanced during its application in HCDI. Consequently, a huge SRC of 33.21 mg g−1 and SRR of 1.50 mg g−1 min−1 were obtained at an operating voltage of 1.2 V in a 500 mg L−1 NaCl solution (Fig. 34b–g).
image file: d4qi03147d-f34.tif
Fig. 34 (a) Synthesis of MoS2/MnO2 composite. (b) Variation in solution conductivity vs. time for MoS2/MnO2, MnO2 and MoS2; (c) SRC vs. time of MoS2/MnO2, MnO2 and MoS2; (d) SRR vs. time of MoS2/MnO2, MnO2 and MoS2; (e) SRC and SRR of MoS2/MnO2 and MnO2 at different potentials; (f) variation in current vs. time and charge efficiency (inset: charge efficiency of MoS2/MnO2 and MnO2); and (g) Kim-Yoon plots for MoS2/MnO2, MnO2 and MoS2. Reproduced from ref. 228 with permission. Copyright 2024, Elsevier.

Besides oxides, hydroxides are also versatile electrode candidates for capacitive applications. Several typical works have been reported on the combination of MoS2 with hydroxides to boost its CDI performance. For example, Zhao et al.229 constructed an FeOOH/Pd/MoS2 hybrid through an ingenious interfacial redox route, which delivered an outstanding desalination capacity and admirable durability. The whole preparation process for the FeOOH/Pd/MoS2 hybrid is presented in Fig. 35a, which includes two-step spontaneous interfacial redox reactions. Initially, MoS2 microflowers are synthesized via a conventional hydrothermal procedure, followed by anchoring Pd nanoparticles through a wet chemical spontaneous interfacial redox reaction route. The driving force for the formation of Pd nanoparticles is based on the reducing property of the DMF solvent. Finally, FeOOH is further deposited onto Pd/MoS2, wherein FeCl2 is utilized as the precursor. The presence of dissolved oxygen caused the oxidation of Fe(III), ensuring the formation of FeOOH. The features of the FeOOH/Pd/MoS2 hybrid resulted in enhanced electrochemical conductivity, and superior capacitive contribution. To evaluate its CDI performance, an HCDI device was assembled using the FeOOH/Pd/MoS2 hybrid as the anode and activated carbon as the cathode (Fig. 35b). As shown in Fig. 35c, the solution conductivity showed a decreasing trend with the working time, indicating efficient CDI capacity. Specially, the desalination capacity of the samples followed the order of MoS2 (26.8 mg g−1) < Pd/MoS2 (33.7 mg g−1) < FeOOH/Pd/MoS2 (41.1 mg g−1). Several factors affecting the performance of the FeOOH/Pd/MoS2//AC cell were further investigated, including initial ion concentration and applied voltage. The result showed that the outlet solution conductivity produced a decreasing trend with a voltage in the range of 0.8 to 1.2 V (Fig. 35d). These phenomena can be ascribed to the high applied voltage causing a stronger electrostatic force, leading to enhanced ion accumulation and intercalation. Significantly, admirable cycling stability and reversibility were also delivered by the FeOOH/Pd/MoS2//AC cell, as shown in Fig. 35e. In addition to single metallic hydroxides, bimetallic layered double hydroxides are also effective to form composites with MoS2, delivering a good CDI performance. Yang et al.18 designed a composite comprised of nickel–ferric-LDH (NiFe-LDH) and MoS2 (NiFe/MoS2), which showed an excellent performance in CDI application for removing high concentrations of Cr(VI). A high deionization capacity of 49.71 mg g−1 was achieved for 100 mg L−1 Cr(VI) and the corresponding removal efficiency was determined to be 99.42%. Significantly, an increased ion selectivity for Cr(VI) in high concentrations of Cl and SO42− was also experimentally verified by the removal mechanism investigation, which involved electrostatic attraction, reduction to low-toxicity Cr(III), and surface complexation. Unfortunately, the experimental removal capacity is still far away from the theoretical value (106.2 mg g−1) and the reusability of LDH-based electrodes in CDI systems needs to be further improved.


image file: d4qi03147d-f35.tif
Fig. 35 (a) Preparation of FeOOH/Pd/MoS2 composite. (b) Schematic of the HCDI device of FeOOH/Pd/MoS2//AC. Dynamic outlet solution conductivity/GACNaCl vs working time plots (c) and working time plots of FeOOH/Pd/MoS2 (d). (e) Cycling performance. Reproduced from ref. 229 with permission. Copyright 2023, Elsevier.

To improve the CDI performance, some ternary composites have also been developed in recent years.230 For example, Huang et al.231 introduced graphene felt (GF) into an MoS2/TiO2-based electrode. In this way, both the wettability and electrochemical characteristics demonstrated an obvious enhancement, leading to a high CDI efficiency and long-term stability. Fig. 36a shows the procedures involved in the preparation of the MoS2/TiO2/GF electrodes. An H3BO3 solution was used to fully penetrate the GF, followed by the deposition of TiO2. Interestingly, the water contact angle measurement indicated a dramatic improvement in the wettability. The pristine GF produced a contact angle of 139.8°, which was much larger than that of TiO2/GF (0°). Finally, MoS2 was further deposited onto TiO2/GF. In this way, the aggregation of MoS2 could be effectively prevented, leading to excellent cycling stability for electrochemical applications. Experimentally, a specific capacitance value of 257.9 F g−1 was obtained for the MoS2/TiO2/GF electrode. In the CDI test, it delivered a desalination efficiency of 34.3% and desalination capacity of 40.96 mg g−1 (Fig. 36b and c). After 50 repeated desalination tests, a desalination retention rate of 92.7% was obtained (Fig. 36d).


image file: d4qi03147d-f36.tif
Fig. 36 (a) Synthesis of MoS2/TiO2/GF composite. Na+ concentration as a function of time for MoS2/TiO2/GF electrode in CDI system with (b) different applied voltages and (c) different NaCl concentrations. (d) SAC and change in conductivity of the synthesized MoS2/TiO2/GF electrodes. Reproduced from ref. 231 with permission. Copyright 2024, Elsevier.

3.5 Composites with other materials

In addition to the above-discussed MoS2-based composites, there are also other functional materials capable of forming composites with MoS2 to deliver boosted CDI performances. As a two-dimensional layered material, transition-metal carbides/nitrides (MXene) are the most promising. MXenes also possess a layered structure and their interlayer interactions are mainly non-covalent bonds, providing great opportunities to modify their properties.232–234 Cai et al.235 designed a 3D MoS2@MXene composite, delivering an excellent CDI performance. As shown in Fig. 37a, the process for the preparation of the MoS2@MXene composite involves two steps. The Ti3C2Tx MXene is obtained by traditional LiF + HCl etching, followed by sonication delamination. After that, MoS2 is deposited onto the surface or the interlayer of MXene to generate the 3D MoS2@MXene composite, which is achieved through a hydrothermal procedure. This 3D MoS2@MXene composite exhibited several benefits. On one hand, the morphology observation and structural identification indicated that a hierarchical 3D structure is formed by MoS2 nanosheets vertically and uniformly anchored on the MXene matrix. This structure is capable of effectively preventing the self-aggregation of MXene flakes and MoS2 nanosheets. On the other hand, more intercalation sites and good electrical conductivity can be offered by the few-layered MXene flakes, ensuring an excellent electrochemical performance. In addition, the water contact angle of MoS2 was determined to be 98.5°, indicating a slightly hydrophobic characteristic. By forming a composite with MXene, the contact angle decreased to 74.4°, showing a hydrophilic feature. Benefiting from these advantages, the heterostructure MoS2@MXene electrode delivered a superior desalination efficiency and charge efficiency compared to the pristine MXene and MoS2 (Fig. 37b). In another work, Chen et al.236 reported a similar process for the synthesis of an MoS2/MXene composite, which demonstrated a superior CDI performance over many reported similar MoS2 based materials.
image file: d4qi03147d-f37.tif
Fig. 37 (a) Procedures for synthesizing MoS2@MXene heterostructure. (b) Desalination capacities and charge efficiencies in different initial concentrations of feed solution. Reproduced from ref. 235 with permission. Copyright 2022, the American Chemical Society.

Considering that both MoS2 and MXene have good CDI desalination performances and strong photocatalysis degradation capacity, it is possible to construct a system integrating CDI and photocatalysis for simultaneously removing ions and organic pollutants from water. Qie et al.237 developed a coupled device, wherein MoS2/Ti3C2Tx was loaded on a carbon fiber surface (CF/MoS2/Ti3C2Tx) to serve as a penetrating electrode (Fig. 38a). Owing to the good combability of layered MoS2 and MXene, the interfacial charge of the lamellar Ti3C2Tx could be well modulated. By introducing a photocatalytic component, this system was capable of simultaneously removing ions and degrading organic pollutants in water. Fig. 38b presents the practical operation process in wastewater treatment, which can be divided into several stages. In stage I, a significant decrease in the conductivity (from 32317.6 to 1212.35 μS cm−1 within 10 s) was observed, indicating an efficient desalination process. In the initial stage, owing to barely all inorganic ions passing through the CDI device, an extremely strong capability of eliminating salt in water will be achieved. Then, with the desalination processing, an increasing trend in the conductivity was observed, which can be attributed to the decrease in DC (stage II). By further prolonging the process, the conductivity of the effluent was observed to maintain almost the same as that of the inflow (stage III). This result can be ascribed to the saturation of DC. To realize the desorption of ions for use in a second cycle, the discharging process is necessary, which was achieved by applying a reverse potential (−0.6 V). In this way, the electrode could completely release all the intercalated ions (stage IV). This release step lasted until the system reached the final equilibrium state (stage V). For practical application, the ability of the coupled system was evaluated to treat wastewater, achieving a yield of 55%. This efficiency was slightly lower than that of the membrane separation technique (near 60%). With an increase in the treatment cycles, the single CF/MoS2/Ti3C2Tx delivered a decreased ion removal capacitance from 49.64 mg g−1 (the first cycle) to 43.73 mg g−1 (the 144th cycle). Furthermore, the corresponding charge efficiency also showed a gradual decreasing trend, from 90.32% (first cycle) to 81.92% (144th cycle) (Fig. 38c). Compared to the simulated experimental results, both the ion removal capacitance and the charge efficiency slightly decreased, which can be attributed to the competition from coexisting ions. Excitingly, this system also enabled the excellent photocatalytic degradation of organic pollutants, which was evidenced by the dramatic decrease in COD, as shown in Fig. 38d. Furthermore, they evidenced that that the hydration radius of ions has a stronger effect on the desalination performance than the ionic valence state. The mutual interference in the CDI desalination and photocatalysis degradation processes was weakened, broadening the application scope of CDI technology.


image file: d4qi03147d-f38.tif
Fig. 38 (a) Schematic of the CF/MoS2/Ti3C2Tx-based four-cell PMCDI device. (b) Conductivity and potential variations in the system during the 72nd cycle. (c) Variations in the ion removal capacity and the charge efficiency during 144 cycles (36 h operation). (d) Comparison of the water quality indexes between the original leachate and treated solution. The removal efficiencies are shown in the inset table. Reproduced from ref. 237 with permission. Copyright 2022, the American Chemical Society.

C3N4 is another representative 2D nanomaterial, showing great promise in energy and environment applications. Similar to MXenes, C3N4 is also capable of loading MoS2 to improve the CDI performance. Tian et al.238 fabricated an MoS2/g-C3N4 composite via a sequential process. As presented in Fig. 39a, they initially prepared C3N4 via the conventional thermal treatment of urea. Afterward, C3N4 was utilized as the substrate to grow MoS2, which was achieved through a hydrothermal process, wherein thiourea served as the sulfur source. The resulting MoS2/g-C3N4 composite exhibited a looser microstructure, in which g-C3N4 nanosheets were coated by MoS2 nanoflowers without obvious agglomeration. Furthermore, the N2 adsorption–desorption measurement revealed the porous structure of the MoS2/g-C3N4 composite. Consequently, numerous channels will be provided for electrolyte transport. The electrochemical investigation indicated that a large specific capacitance of 118.3 F g−1 was delivered by the MoS2/g-C3N4 composite electrode at 1 A g−1. When the current density increased to 10 A g−1, the capacitance remained at 61.7 F g−1, demonstrating a remarkable rate capability. Benefiting from these characteristics, an excellent CDI efficiency could be obtained when using the MoS2/g-C3N4-based electrode. As presented in Fig. 39b, the conductivity of the MoS2/g-C3N4 composite-based CDI cell showed a dramatic decreasing trend from the initial to saturation state, ranging from 500 to 372.67 μS cm−1. This result indicates that the MoS2/g-C3N4 composite-based electrode delivered an excellent CDI performance. By contrast, the pristine g-C3N4 and MoS2-based electrodes showed a slight decline in conductivity, ranging from 500 to 467.15 and 421.46 μS cm−1, respectively. According to the decrease in conductivity, the SAC values were calculated to be 6.25, 14.91, and 24.18 mg g−1 for the g-C3N4, MoS2, and MoS2/g-C3N4 composite-based electrodes, respectively. The enhanced CDI performance of the MoS2/g-C3N4 composite-based electrode can be attributed to the intimate contact between MoS2 and the 2D g-C3N4 nanosheets and its controllable porous nanostructure. The stability tests of the MoS2/g-C3N4 composite electrode demonstrated its good stability in consecutive electrosorption–desorption cycling (Fig. 39c).


image file: d4qi03147d-f39.tif
Fig. 39 (a) Procedure for synthesizing MoS2/g-C3N4 composite. (b) CDI performance comparison of g-C3N4, MoS2, and MoS2/g-C3N4-based electrodes. (c) Cycling stability of MoS2/g-C3N4-based electrode. Reproduced from ref. 238 with permission. Copyright 2020, Elsevier.

The improvement in CDI performance by the combination of MoS2 and g-C3N4 can be ascribed to several factors. On one hand, g-C3N4 possesses a rich nitrogen content, unique electronic structure, chemical and thermal stability, and environmentally acceptable character, capable of providing multifunctional properties in catalysis and energy conversion. However, its low intrinsic electronic conductivity is the main limitation in its electrochemical application. Thus, the integration of MoS2 and g-C3N4 may bring several benefits. On one hand, their electrochemical properties can be significantly improved, originating from the synergistic effects between these two components. On the other hand, the re-stacking of MoS2 can be dramatically prevented, improving its stability.

4 Conclusion and perspectives

The intrinsic features such as straightforward operation, low energy consumption, and no secondary pollution of CDI makes it a promising method for decreasing the ionic concentration of salty water and heavy metal wastewater treatment. MoS2 is a fascinating 2D layered material exhibiting various physicochemical properties, making it suitable for designing high-performance electrochemical devices, particularly CDI devices. In this review, we summarized and discussed different techniques for the synthesis of MoS2, focusing on the research developments for CDI application. Many MoS2-based CDI devices have been investigated for brackish water desalination and wastewater deionization but issues still exist, which can be outlined as follows:

(1) Safe and efficient methods need to be developed for the large-scale preparation of MoS2. Currently, the most used top-down method for the preparation of MoS2 involves various exfoliation processes, which suffer from low efficiency and poor control of the thickness and number of layers of MoS2 nanoflakes (or nanosheets). In the case of bottom-up methods (such as chemical vapor deposition, colloidal synthesis, hydrothermal/solvothermal synthesis, and salt-assisted synthesis), they involve the use of high-toxicity precursors, such as KSCN and metalorganic compounds, or generate harmful waste, such as H2S and sulfur oxides. In addition, some of them consume plenty of energy or solvents. Consequently, other processes using less hazardous etchants that are environmentally friendly and energy efficient should be developed.

(2) More advanced electrode materials are urgently needed. According to the studies discussed in Section 3, one can easily find that most of the current works are focused on developing MoS2/carbon composites for gaining high-performance CDI cells, in addition to adjusting the structures/defects in pristine MoS2. In comparison, there are much less works on composite comprised of MoS2 and other species (conducting polymers, other inorganic materials, MXenes, etc.).

(3) Ion selectivity should be considered. Considering sustainable development, both brackish water and ion-containing wastewater have environmental/resource impacts, and thus scientists are actively developing advanced CDI electrode materials for capturing target ions. In the case of MoS2-based CDI cells, more effort should be devoted to modifying MoS2-based CDI electrodes, through designing particular micro/nanostructures, functionalizing specific groups or compositing with other species to improve the CDI selectivity towards the target ions and boost their efficiency in water treatment applications.

(4) To realize the above-mentioned goals, the incorporation of current advanced technologies such as artificial intelligence (AI), big data analysis and machine learning into CDI systems to design and guide the preparation of MoS2-based electrodes and fully realize the selective elimination of ions from water. AI-driven CDI is expected to make positive contributions to addressing the global fresh water shortage and pollution challenges by dynamically optimizing the ion selectivity in real-time and adapting to changing ion concentrations and water compositions.

For future specific studies, there are at least three directions, as follows:

(1) Phase engineering. More attention should be paid to controlling the phase structure of MoS2 and tuning its properties. It is well known that the 1T phase exhibits superior electronic conductivity but inferior stability compared to the 2H phase. Thus, it is necessary to balance these two aspects, which requires scientists to devote more effort to the mechanism of synthesis processes.

(2) Electrode–electrolyte interfaces. The interfaces between the electrode and electrolyte play a significant role in determining the overall performance of electrochemical devices. However, to date, direct techniques to reveal the processes occurring at the interfaces are still lacking. The accumulation of ions at the electrode/electrolyte interfaces may be affected by many factors, such as the wettability of the electrode and the ionic migration characteristics in the system. Thus, to achieve this goal, more in situ characterization techniques should be developed in the future.

(3) The combination of CDI and other techniques is necessary. It is well known that in real wastewater and seawater, various species are simultaneously present besides charged ions, such as organic pollutants. Thus, to obtain high-quality fresh water, simultaneously removing these contaminants by incorporating photocatalytic function into CDI cells is preferred. Expectedly, more and more functions will be integrated into MXene-based CDI systems in the future.

In conclusion, benefiting from the advantages of MoS2 including unique layered structures and ease of modification, more and more MoS2-based advanced materials with unique structures and multifunction can be developed. Therefore, it is expected that that these materials will dramatically promote the development of CDI devices.

Data availability

No primary research results, software or code have been included and no new data were generated or analyzed as part of this review.

Conflicts of interest

The authors declare no competing interest.

References

  1. Y. Liu, L. Wang, Q. Ma, X. Xu, X. Gao, H. Zhu, T. Feng, X. Dou, M. Eguchi, Y. Yamauchi and X. Yuan, Simultaneous generation of residue-free reactive oxygen species and bacteria capture for efficient electrochemical water disinfection, Nat. Commun., 2024, 15, 10175 CrossRef CAS PubMed.
  2. Y. Liu, Y. Zhang, Y. Zhang, Q. Zhang, X. Gao, X. Dou, H. Zhu, X. Yuan and L. Pan, MoC nanoparticle-embedded carbon nanofiber aerogels as flow-through electrodes for highly efficient pseudocapacitive deionization, J. Mater. Chem. A, 2020, 8, 1443–1450 RSC.
  3. S. He, T. Zhu, Y. Wang, W. Xiong, X. Gao and E. Zhang, Application of capacitive deionization technology in water treatment and coupling technology: a review, Environ. Sci.: Water Res. Technol., 2024, 10, 2313–2340 RSC.
  4. W. Zhai, Z. Li, Y. Wang, L. Zhai, Y. Yao, S. Li, L. Wang, H. Yang, B. Chi, J. Liang, Z. Shi, Y. Ge, Z. Lai, Q. Yun, A. Zhang, Z. Wu, Q. He, B. Chen, Z. Huang and H. Zhang, Phase Engineering of Nanomaterials: Transition Metal Dichalcogenides, Chem. Rev., 2024, 124(7), 4479–4539 CrossRef.
  5. S. Singh, A. Modak, K. K. Pant, A. Sinhamahapatra and P. Biswas, MoS2–Nanosheets-Based Catalysts for Photocatalytic CO2 Reduction: A Review, ACS Appl. Nano Mater., 2021, 4(9), 8644–8667 CrossRef.
  6. J. Mao, J. He, W. F. Io, F. Guo, Z. Wu, M. Yang and J. Hao, Strain-Engineered Ferroelectricity in 2H Bilayer MoS2, ACS Nano, 2024, 18(44), 30360–30367 CrossRef PubMed.
  7. A. Giri, G. Park and U. Jeong, Layer-Structured Anisotropic Metal Chalcogenides: Recent Advances in Synthesis, Modulation, and Applications, Chem. Rev., 2023, 123(7), 3329–3442 CrossRef PubMed.
  8. J. Strachan, A. F. Masters and T. Maschmeyer, 3R-MoS2 in Review: History, Status, and Outlook, ACS Appl. Energy Mater., 2021, 4(8), 7405–7418 CrossRef CAS.
  9. J. Li, X. Li, L. He, H. Guo, W. Xia, B. Sun, C. Cao, L. Sha and D. Zhou, MoS2-Based Nanocomposites for Microwave Absorption: A Review, ACS Appl. Nano Mater., 2024, 7(6), 5761–5775 CrossRef CAS.
  10. J. Liang, H. Li, L. Chen, M. Ren, O. A. Fakayode, J. Han and C. Zhou, Efficient hydrogen evolution reaction performance using lignin-assisted chestnut shell carbon-loaded molybdenum disulfide, Ind. Crops Prod., 2023, 193, 116214 CrossRef CAS.
  11. N. Liang, X. Hu, W. Li, Y. Wang, Z. Guo, X. Huang, Z. Li, X. Zhang, J. Zhang, J. Xiao, X. Zou and J. Shi, A dual-signal fluorescent sensor based on MoS2 and CdTe quantum dots for tetracycline detection in milk, Food Chem., 2022, 378, 132076 CrossRef CAS.
  12. Q. Ji, X. Yu, L. Chen, O. P. N. Yarley and C. Zhou, Facile preparation of sugarcane bagasse-derived carbon supported MoS2 nanosheets for hydrogen evolution reaction, Ind. Crops Prod., 2021, 172, 114064 CrossRef CAS.
  13. P. Chavalekvirat, W. Hirunpinyopas, K. Deshsorn, K. Jitapunkul and P. Iamprasertkun, Liquid Phase Exfoliation of 2D Materials and Its Electrochemical Applications in the Data-Driven Future, Precis. Chem., 2024, 2(7), 300–329 CrossRef CAS.
  14. F. Xing, T. Li, J. Li, H. Zhu, N. Wang and X. Cao, Chemically exfoliated MoS2 for capacitive deionization of saline water, Nano Energy, 2017, 31, 590–595 CrossRef.
  15. M. Zhang, F. Jia, M. Dai and S. Song, Combined electrosorption and chemisorption of low concentration Pb(II) from aqueous solutions with molybdenum disulfide as electrode, Appl. Surf. Sci., 2018, 455, 258–266 CrossRef.
  16. B. Pan, Y. Wang, H. Li, W. Yi and Y. Pan, Preparation and Electrosorption Desalination Performance of Peanut Shell-Based Activated Carbon and MoS2, Int. J. Electrochem. Sci., 2020, 15, 1861–1880 CrossRef.
  17. Z. Zhao, J. Zhao, Y. Sun, M. Ye and X. Wen, In situ construction of 3D hierarchical MoS2/CoS2@TiO2 nanotube hybrid electrodes with superior capacitive performance toward water treatment, Chem. Eng. J., 2022, 429, 132582 CrossRef.
  18. D. Yang, X. Li, Y. Li, W. Song, T. Yan, Y. Cui and L. Yan, Capacitive deionization of high concentrations of hexavalent chromium using nickel–ferric-layered double hydroxide/molybdenum disulfide asymmetric electrode, J. Colloid Interface Sci., 2023, 634, 793–803 CrossRef PubMed.
  19. J. Wang, Z. Chen, R. Yuan, J. Luo, B. Zhang, K. Ji, M. Li, J. Xiao and K. Sun, Innovative dual-mode device integrating capacitive desalination and solar vapor generation for high-efficiency seawater desalination, J. Energy Chem., 2025, 100, 171–179 CrossRef.
  20. R. Aggarwal, D. Saini, R. Mitra, S. K. Sonkar, A. K. Sonker and G. Westman, From Bulk Molybdenum Disulfide (MoS2) to Suspensions of Exfoliated MoS2 in an Aqueous Medium and Their Applications, Langmuir, 2024, 40(19), 9855–9872 CrossRef PubMed.
  21. Y. Jia, Y. Zhang, H. Xu, J. Li, M. Gao and X. Yang, Recent Advances in Doping Strategies to Improve Electrocatalytic Hydrogen Evolution Performance of Molybdenum Disulfide, ACS Catal., 2024, 14(7), 4601–4637 CrossRef CAS.
  22. Q. Dong, Q. Wei and Z. Tang, Molybdenum Disulfide (MoS2): An Emerging Multifunctional Nanomaterial for Sensing and Removal of Agricultural Contaminants, ACS Agric. Sci. Technol., 2024, 4(2), 173–192 CrossRef CAS.
  23. Y. Yu, L. Lu, Q. Yang, A. Zupanic, Q. Xu and L. Jiang, Using MoS2 Nanomaterials to Generate or Remove Reactive Oxygen Species: A Review, ACS Appl. Nano Mater., 2021, 4(8), 7523–7537 CrossRef CAS.
  24. Z. Wang and B. Mi, Environmental Applications of 2D Molybdenum Disulfide (MoS2) Nanosheets, Environ. Sci. Technol., 2017, 51(15), 8229–8244 CrossRef CAS PubMed.
  25. E. Singh, P. Singh, K. S. Kim, G. Y. Yeom and H. S. Nalwa, Flexible Molybdenum Disulfide (MoS2) Atomic Layers for Wearable Electronics and Optoelectronics, ACS Appl. Mater. Interfaces, 2019, 11(12), 11061–11105 CrossRef CAS.
  26. Q. Li, D. Liang, D. Wang, L. Ling, L. Jiang, F. Qiu and M. Qian, Recent Progress and Outlook on the Emerging Low-Dimensional MoS2 Nanostructures for Microwave Absorption, ACS Appl. Electron. Mater., 2024, 6(1), 120–143 CrossRef CAS.
  27. T. P. Nguyen, S. Choi, J.-M. Jeon, K. C. Kwon, H. W. Jang and S. Y. Kim, Transition Metal Disulfide Nanosheets Synthesized by Facile Sonication Method for the Hydrogen Evolution Reaction, J. Phys. Chem. C, 2016, 120(7), 3929–3935 CrossRef CAS.
  28. M. Mohiuddin, Y. Wang, A. Zavabeti, N. Syed, R. S. Datta, H. Ahmed, T. Daeneke, S. P. Russo, A. R. Rezk, L. Y. Yeo and K. Kalantar-Zadeh, Liquid Phase Acoustic Wave Exfoliation of Layered MoS2: Critical Impact of Electric Field in Efficiency, Chem. Mater., 2018, 30(16), 5593–5601 CrossRef CAS.
  29. E. P. Nguyen, B. J. Carey, T. Daeneke, J. Z. Ou, K. Latham, S. Zhuiykov and K. Kalantar-zadeh, Investigation of Two-Solvent Grinding-Assisted Liquid Phase Exfoliation of Layered MoS2, Chem. Mater., 2015, 27(1), 53–59 CrossRef CAS.
  30. A. Gupta, V. Arunachalam and S. Vasudevan, Liquid-Phase Exfoliation of MoS2 Nanosheets: The Critical Role of Trace Water, J. Phys. Chem. Lett., 2016, 7(23), 4884–4890 CrossRef CAS PubMed.
  31. P. Chavalekvirat, P. Nakkiew, T. Kunaneksin, W. Hirunpinyopas, W. Busayaporn and P. Iamprasertkun, Tuning Surface Energy to Enhance MoS2 Nanosheet Production via Liquid-Phase Exfoliation: Understanding the Electrochemical Adsorption of Cesium Chloride, Inorg. Chem., 2023, 62(32), 12851–12861 CrossRef CAS.
  32. P. Wang, J. Wu, X. Xiao, Y. Fan, X. Han and Y. Sun, Engineering Injectable Coassembled Hydrogel by Photothermal Driven Chitosan-Stabilized MoS2 Nanosheets for Infected Wound Healing, ACS Nano, 2024, 18(39), 26961–26974 CrossRef CAS.
  33. G. S. Bang, K. W. Nam, J. Y. Kim, J. Shin, J. W. Choi and S.-Y. Choi, Effective Liquid-Phase Exfoliation and Sodium Ion Battery Application of MoS2 Nanosheets, ACS Appl. Mater. Interfaces, 2014, 6(10), 7084–7089 CrossRef CAS.
  34. Y. Li, X. Yin and W. Wu, Preparation of Few-Layer MoS2 Nanosheets via an Efficient Shearing Exfoliation Method, Ind. Eng. Chem. Res., 2018, 57(8), 2838–2846 CrossRef CAS.
  35. X. Yu, M. S. Prévot and K. Sivula, Multiflake Thin Film Electronic Devices of Solution Processed 2D MoS2 Enabled by Sonopolymer Assisted Exfoliation and Surface Modification, Chem. Mater., 2014, 26(20), 5892–5899 CrossRef CAS.
  36. E. Varrla, C. Backes, K. R. Paton, A. Harvey, Z. Gholamvand, J. McCauley and J. N. Coleman, Large-Scale Production of Size-Controlled MoS2 Nanosheets by Shear Exfoliation, Chem. Mater., 2015, 27(3), 1129–1139 CrossRef CAS.
  37. A. Gupta and S. Vasudevan, Understanding Surfactant Stabilization of MoS2 Nanosheets in Aqueous Dispersions from Zeta Potential Measurements and Molecular Dynamics Simulations, J. Phys. Chem. C, 2018, 122(33), 19243–19250 CrossRef CAS.
  38. A. Gupta, V. Arunachalam and S. Vasudevan, Water Dispersible Positively and Negatively Charged MoS2 Nanosheets: Surface Chemistry and the Role of Surfactant Binding, J. Phys. Chem. Lett., 2015, 6(4), 739–744 CrossRef CAS PubMed.
  39. N. K. Oh, H. J. Lee, K. Choi, J. Seo, U. Kim, J. Lee, Y. Choi, S. Jung, J. H. Lee, H. S. Shin and H. Park, Nafion-Mediated Liquid-Phase Exfoliation of Transition Metal Dichalcogenides and Direct Application in Hydrogen Evolution Reaction, Chem. Mater., 2018, 30(14), 4658–4666 CrossRef CAS.
  40. S. Karunakaran, S. Pandit, B. Basu and M. De, Simultaneous Exfoliation and Functionalization of 2H-MoS2 by Thiolated Surfactants: Applications in Enhanced Antibacterial Activity, J. Am. Chem. Soc., 2018, 140(39), 12634–12644 CrossRef CAS.
  41. A. Z. Melaku, W.-T. Chuang, C.-W. Chiu, J.-Y. Lai and C.-C. Cheng, Controlling the Hierarchical Structures of Molybdenum Disulfide Nanomaterials via Self-Assembly of Supramolecular Polymers in Water, Chem. Mater., 2022, 34(7), 3333–3345 CrossRef CAS.
  42. L. Zong, M. Li and C. Li, Bioinspired Coupling of Inorganic Layered Nanomaterialswith Marine Polysaccharides for Efficient Aqueous Exfoliation and Smart Actuating Hybrids, Adv. Mater., 2017, 29, 1604691 CrossRef.
  43. G. Guan, S. Zhang, S. Liu, Y. Cai, M. Low, C. P. Teng, I. Y. Phang, Y. Cheng, K. L. Duei, B. M. Srinivasan, Y. Zheng, Y.-W. Zhang and M.-Y. Han, Protein Induces Layer-by-Layer Exfoliation of Transition Metal Dichalcogenides, J. Am. Chem. Soc., 2015, 137(19), 6152–6155 CrossRef CAS.
  44. S. Roy, P. Haloi, R. Choudhary, S. Chawla, M. Kumari, V. Badireenath Konkimalla and A. Jaiswal, Quaternary Pullulan-Functionalized 2D MoS2 Glycosheets: A Potent Bactericidal Nanoplatform for Efficient Wound Disinfection and Healing, ACS Appl. Mater. Interfaces, 2023, 15(20), 24209–24227 CrossRef CAS.
  45. X. Xu, J. Wu, Z. Meng, Y. Li, Q. Huang, Y. Qi, Y. Liu, D. Zhan and X. Y. Liu, Enhanced Exfoliation of Biocompatible MoS2 Nanosheets by Wool Keratin, ACS Appl. Nano Mater., 2018, 1(10), 5460–5469 CrossRef CAS.
  46. C. Zhang, D.-F. Hu, J.-W. Xu, M.-Q. Ma, H. Xing, K. Yao, J. Ji and Z.-K. Xu, Polyphenol-Assisted Exfoliation of Transition Metal Dichalcogenides into Nanosheets as Photothermal Nanocarriers for Enhanced Antibiofilm Activity, ACS Nano, 2018, 12(12), 12347–12356 CrossRef CAS.
  47. I.-W. P. Chen, Y.-H. Li, Y.-L. Su and G.-X. Huang, Layer-by-Layer Exfoliation of Transition-Metal Dichalcogenides by Amino Acid in Water for Promoting Hydrogen Evolution Reaction, J. Phys. Chem. C, 2022, 126(14), 6207–6214 CrossRef CAS.
  48. A. K. Sonker, S. Xiong, R. Aggarwal, M. Olsson, A. Spule, S. Hosseini, S. K. Sonkar, A. Matic and G. Westman, Exfoliated MoS2 Nanosheet/Cellulose Nanocrystal Flexible Composite Films as Electrodes for Zinc Batteries, ACS Appl. Nano Mater., 2023, 6(10), 8270–8278 CrossRef CAS.
  49. S. Cao, L. Shi, M. Miao, J. Fang, H. Zhao and X. Feng, Solution-processed flexible paper-electrode for lithium-ion batteries based on MoS2 nanosheets exfoliated with cellulose nanofibrils, Electrochim. Acta, 2019, 298, 22–30 CrossRef CAS.
  50. P. Yadav, A. Das, M. K. Sharma, S. Ash and A. K. Ganguli, A biodegradable polymer-assisted efficient and universal exfoliation route to a stable few layer dispersion of transition metal dichalcogenides, Mater. Chem. Phys., 2022, 276, 125347 CrossRef CAS.
  51. W.-S. He, L. Zhao, H. Yang, J. Rui, J. Li and Z.-Y. Chen, Novel Synthesis of Phytosterol Ferulate Using Acidic Ionic Liquids as a Catalyst and Its Hypolipidemic Activity, J. Agric. Food Chem., 2024, 72(4), 2309–2320 CrossRef CAS.
  52. K. Salikolimi, A. A. Sudhakar and Y. Ishida, Functional Ionic Liquid Crystals, Langmuir, 2020, 36(40), 11702–11731 CrossRef CAS.
  53. M. Zhu, Ionic-liquid/metal-organic-framework composites: Synthesis and emerging sustainable applications, Inorg. Chem. Front., 2024, 12, 39–84 RSC.
  54. I. D. Boateng, D. A. Soetanto, F. Li, X.-M. Yang and Y.-Y. Li, Separation and purification of polyprenols from Ginkgo biloba L. leaves by bulk ionic liquid membrane and optimizing parameters, Ind. Crops Prod., 2021, 170, 113828 CrossRef CAS.
  55. A. Nezhadali, R. Mohammadi and M. Mojarrab, An overview on pollutants removal from aqueous solutions via bulk liquid membranes (BLMs): Parameters that influence the effectiveness, selectivity and transport kinetic, J. Environ. Chem. Eng., 2019, 7, 103339 CrossRef CAS.
  56. I. D. Boateng, X.-M. Yang, H. Yin and W. Liu, Separation and purification of polyprenols from Ginkgo biloba leaves by silver ion anchored on imidazole-based ionic liquid functionalized mesoporous MCM-41 sorbent, Food Chem., 2024, 450, 139284 CrossRef CAS PubMed.
  57. L. Zhang, Y. Hu, X. Wang, A. Zhang, X. Gao, A. E.-G. A. Yagoub, H. Ma and C. Zhou, Ultrasound-Assisted Synthesis of Potentially Food-Grade Nano-Zinc Oxide in Ionic Liquids: A Safe, Green, Efficient Approach and Its Acoustics Mechanism, Foods, 2022, 11(11), 1656 CrossRef CAS PubMed.
  58. R. Tian, X. Jia, M. Lan, S. Wang, Y. Li, J. Yang, D. Shao, L. Feng, Q. Su and H. Song, Ionic Liquid Crystals Confining Ultrathin MoS2 Nanosheets: A High-Concentration and Stable Aqueous Dispersion, ACS Sustainable Chem. Eng., 2022, 10(13), 4186–4197 CrossRef CAS.
  59. W. Guo, M. Li, X. Wu, Y. Liu, T. Ou, C. Xiao, Z. Qiu, Y. Zheng and Y. Wang, Nonvolatile n-Type Doping and Metallic State in Multilayer-MoS2 Induced by Hydrogenation Using Ionic-Liquid Gating, Nano Lett., 2022, 22(22), 8957–8965 CrossRef CAS.
  60. F. Wang, P. Stepanov, M. Gray, C. N. Lau, M. E. Itkis and R. C. Haddon, Ionic Liquid Gating of Suspended MoS2 Field Effect Transistor Devices, Nano Lett., 2015, 15(8), 5284–5288 CrossRef CAS.
  61. Y. Lee, J. Lee, S. Kim and H. S. Park, Rendering High Charge Density of States in Ionic Liquid-Gated MoS2 Transistors, J. Phys. Chem. C, 2014, 118(31), 18278–18282 CrossRef CAS.
  62. Q. Ji, X. Yu, P. Wu, A. E.-G. A. Yagoub, L. Chen, A. T. Mustapha and C. Zhou, Pretreatment of sugarcane bagasse with deep eutectic solvents affect the structure and morphology of lignin, Ind. Crops Prod., 2021, 173, 114108 CrossRef CAS.
  63. H. Li, J. Liang, L. Chen, M. Ren and C. Zhou, Utilization of walnut shell by deep eutectic solvents: enzymatic digestion of cellulose and preparation of lignin nanoparticles, Ind. Crops Prod., 2023, 192, 116034 CrossRef CAS.
  64. E. L. Smith, A. P. Abbott and K. S. Ryder, Deep Eutectic Solvents (DESs) and Their Applications, Chem. Rev., 2014, 114(21), 11060–11082 CrossRef CAS PubMed.
  65. W. Huang, X. He, J. Wu, X. Ma, J. Han, L. Wang and Y. Wang, The evaluation of deep eutectic solvents and ionic liquids as cosolvents system for improving cellulase properties, Ind. Crops Prod., 2023, 197, 116555 CrossRef CAS.
  66. Y. Ma, Y. Yang, T. Li, S. Hussain and M. Zhu, Deep eutectic solvents as an emerging green platform for the synthesis of functional materials, Green Chem., 2024, 26, 3627–3669 RSC.
  67. Q. Ji, X. Yu, A. E.-G. A. Yagoub, L. Chen and C. Zhou, Efficient removal of lignin from vegetable wastes by ultrasonic and microwave-assisted treatment with ternary deep eutectic solvent, Ind. Crops Prod., 2020, 149, 112357 CrossRef CAS.
  68. Q. Ma, Q. Ji, L. Chen, Z. Zhu, S. Tu, C. E. Okonkwo, P. Out and C. Zhou, Multimode ultrasound and ternary deep eutectic solvent sequential pretreatments enhanced the enzymatic saccharification of corncob biomass, Ind. Crops Prod., 2022, 188, 115574 CrossRef CAS.
  69. W. Ashraf, A. Rehman, A. Hussain, A. Karim, H. R. Sharif, M. Siddiquy and Z. Lianfu, Optimization of Extraction Process and Estimation of Flavonoids from Fenugreek Using Green Extracting Deep Eutectic Solvents Coupled with Ultrasonication, Food Bioprocess Technol., 2024, 17, 887–903 CrossRef CAS.
  70. L. Zeb, A. S. Gerhardt, B. A. Johannesen, J. Underhaug and M. Jordheim, Ultrasonic-Assisted Water-Rich Natural Deep Eutectic Solvents for Sustainable Polyphenol Extraction from Seaweed: A Case Study on Cultivated Saccharina latissimi, ACS Sustainable Chem. Eng., 2024, 12(40), 14921–14929 CrossRef CAS.
  71. M. Zhou, O. A. Fakayode, M. Ren, H. Li, J. Liang and C. Zhou, Green and sustainable extraction of lignin by deep eutectic solvent, its antioxidant activity, and applications in the food industry, Crit. Rev. Food Sci. Nutr., 2024, 64, 7201–7219 CrossRef CAS PubMed.
  72. M. Rezaee, F. Feyzi and S. Javanshir, Application of Response Surface Methodology for Selective Extraction of Lithium Using a Hydrophobic Deep Eutectic Solvent, Ind. Eng. Chem. Res., 2025, 64(4), 2294–2308 CrossRef CAS.
  73. R. S. Lamarca, S. D. S Ferreira, M. P. Abuçafy, L. S. Silva, K. A. Pishro, C. D. B. Amaral, J. O. Fernandes, S. C. Cunha, D. F. Franco and M. H. Gonzalez, Advances in Sustainable Extraction of Arsenic and Cadmium from Sewage Sludge: Use of Hydrophobic Deep Eutectic Solvent Modified with Magnetic Nanoparticles, ACS Sustainable Chem. Eng., 2024, 12(39), 14371–14379 CrossRef CAS.
  74. Z. Mohammadpour, S. H. Abdollahi and A. Safavi, Sugar-Based Natural Deep Eutectic Mixtures as Green Intercalating Solvents for High-Yield Preparation of Stable MoS2 Nanosheets: Application to Electrocatalysis of Hydrogen Evolution Reaction, ACS Appl. Energy Mater., 2018, 1(11), 5896–5906 CrossRef CAS.
  75. Y. Biswas, M. Dule and T. K. Mandal, Poly(ionic liquid)-Promoted Solvent-Borne Efficient Exfoliation of MoS2/MoSe2 Nanosheets for Dual-Responsive Dispersion and Polymer Nanocomposites, J. Phys. Chem. C, 2017, 121(8), 4747–4759 CrossRef CAS.
  76. M. Zhu and Y. Yang, Poly(ionic liquid)s: an emerging platform for green chemistry, Green Chem., 2024, 26, 5022–5102 RSC.
  77. X. Wang and P. Wu, Aqueous Phase Exfoliation of Two-Dimensional Materials Assisted by Thermoresponsive Polymeric Ionic Liquid and Their Applications in Stimuli-Responsive Hydrogels and Highly Thermally Conductive Films, ACS Appl. Mater. Interfaces, 2018, 10(3), 2504–2514 CrossRef CAS.
  78. D. Gopalakrishnan, D. Damien and M. M. Shaijumon, MoS2 Quantum Dot-Interspersed Exfoliated MoS2 Nanosheets, ACS Nano, 2014, 8(5), 5297–5303 CrossRef CAS.
  79. W. Dai, H. Dong, B. Fugetsu, Y. Cao, H. Lu, X. Ma and X. Zhang, Tunable Fabrication of Molybdenum Disulfide Quantum Dots for Intracellular MicroRNA Detection and Multiphoton Bioimaging, Small, 2015, 11, 4158–4164 CrossRef CAS.
  80. S. Xu, D. Li and P. Wu, One-Pot, Facile, and Versatile Synthesis of Monolayer MoS2/WS2 Quantum Dots as Bioimaging Probes and Efficient Electrocatalysts for Hydrogen Evolution Reaction, Adv. Funct. Mater., 2015, 25, 1127–1136 CrossRef CAS.
  81. X. S. Chu, A. Yousaf, D. O. Li, A. A. Tang, A. Debnath, D. Ma, A. A. Green, E. J. G. Santos and Q. H. Wang, Direct Covalent Chemical Functionalization of Unmodified Two-Dimensional Molybdenum Disulfide, Chem. Mater., 2018, 30(6), 2112–2128 CrossRef CAS.
  82. M. Vera-Hidalgo, E. Giovanelli, C. Navío and E. M. Pérez, Mild Covalent Functionalization of Transition Metal Dichalcogenides with Maleimides: A “Click” Reaction for 2H-MoS2 and WS2, J. Am. Chem. Soc., 2019, 141(9), 3767–3771 CrossRef CAS PubMed.
  83. N. Li, Z. Liu, M. Liu, C. Xue, Q. Chang, H. Wang, Y. Li, Z. Song and S. Hu, Facile Synthesis of Carbon Dots@2D MoS2 Heterostructure with Enhanced Photocatalytic Properties, Inorg. Chem., 2019, 58(9), 5746–5752 CrossRef CAS PubMed.
  84. K. C. Knirsch, N. C. Berner, H. C. Nerl, C. S. Cucinotta, Z. Gholamvand, N. McEvoy, Z. Wang, I. Abramovic, P. Vecera, M. Halik, S. Sanvito, G. S. Duesberg, V. Nicolosi, F. Hauke, A. Hirsch, J. N. Coleman and C. Backes, Basal-Plane Functionalization of Chemically Exfoliated Molybdenum Disulfide by Diazonium Salts, ACS Nano, 2015, 9(6), 6018–6030 CrossRef CAS.
  85. J. Zheng, H. Zhang, S. Dong, Y. Liu, C. T. Nai, H. S. Shin, H. Y. Jeong, B. Liu and K. P. Loh, High yield exfoliation of two-dimensional chalcogenides using sodium naphthalenide, Nat. Commun., 2014, 5, 2995 CrossRef PubMed.
  86. X. Zhu, Z. Su, C. Wu, H. Cong, X. Ai, H. Yang and J. Qian, Exfoliation of MoS2 Nanosheets Enabled by a Redox-Potential-Matched Chemical Lithiation Reaction, Nano Lett., 2022, 22(7), 2956–2963 CrossRef CAS PubMed.
  87. J. V. Pondick, A. Kumar, M. Wang, S. Yazdani, J. M. Woods, D. Y. Qiu and J. J. Cha, Heterointerface Control over Lithium-Induced Phase Transitions in MoS2 Nanosheets: Implications for Nanoscaled Energy Materials, ACS Appl. Nano Mater., 2021, 4(12), 14105–14114 CrossRef CAS.
  88. J. Zhang, T. Ji, H. Jin, Z. Wang, M. Zhao, D. He, G. Luo and B. Mao, Mild Liquid-Phase Exfoliation of Transition Metal Dichalcogenide Nanosheets for Hydrogen Evolution, ACS Appl. Nano Mater., 2022, 5(6), 8020–8028 CrossRef CAS.
  89. X. Fan, P. Xu, D. Zhou, Y. Sun, Y. C. Li, M. A. T. Nguyen, M. Terrones and T. E. Mallouk, Fast and Efficient Preparation of Exfoliated 2H MoS2 Nanosheets by Sonication-Assisted Lithium Intercalation and Infrared Laser-Induced 1T to 2H Phase Reversion, Nano Lett., 2015, 15(9), 5956–5960 CrossRef CAS.
  90. X. Fan, P. Xu, Y. C. Li, D. Zhou, Y. Sun, M. A. T. Nguyen, M. Terrones and T. E. Mallouk, Controlled Exfoliation of MoS2 Crystals into Trilayer Nanosheets, J. Am. Chem. Soc., 2016, 138(15), 5143–5149 CrossRef CAS.
  91. N. Liu, P. Kim, J. H. Kim, J. H. Ye, S. Kim and C. J. Lee, Large-Area Atomically Thin MoS2 Nanosheets Prepared Using Electrochemical Exfoliation, ACS Nano, 2014, 8(7), 6902–6910 CrossRef CAS PubMed.
  92. S. García-Dalí, J. I. Paredes, J. M. Munuera, S. Villar-Rodil, A. Adawy, A. Martínez-Alonso and J. M. D. Tascón, Aqueous Cathodic Exfoliation Strategy toward Solution-Processable and Phase-Preserved MoS2 Nanosheets for Energy Storage and Catalytic Applications, ACS Appl. Mater. Interfaces, 2019, 11(40), 36991–37003 CrossRef.
  93. S. García-Dalí, J. I. Paredes, S. Villar-Rodil, A. Martínez-Jódar, A. Martínez-Alonso and J. M. D. Tascón, Molecular Functionalization of 2H-Phase MoS2 Nanosheets via an Electrolytic Route for Enhanced Catalytic Performance, ACS Appl. Mater. Interfaces, 2021, 13(28), 33157–33171 CrossRef.
  94. Y. Zhuo, I. A. Kinloch and M. A. Bissett, Simultaneous Electrochemical Exfoliation and Functionalization of 2H-MoS2 for Supercapacitor Electrodes, ACS Appl. Nano Mater., 2023, 6(19), 18062–18070 CrossRef CAS.
  95. P. Zuo, L. Jiang, X. Li, M. Tian, L. Ma, C. Xu, Y. Yuan, X. Li, X. Chen and M. Liang, Phase-Reversed MoS2 Nanosheets Prepared through Femtosecond Laser Exfoliation and Chemical Doping, J. Phys. Chem. C, 2021, 125(15), 8304–8313 CrossRef.
  96. W. Wu, J. Xu, X. Tang, P. Xie, X. Liu, J. Xu, H. Zhou, D. Zhang and T. Fan, Two-Dimensional Nanosheets by Rapid and Efficient Microwave Exfoliation of Layered Materials, Chem. Mater., 2018, 30(17), 5932–5940 CrossRef.
  97. V.-T. Nguyen, T.-Y. Yang, P. A. Le, P.-J. Yen, Y.-L. Chueh and K.-H. Wei, New Simultaneous Exfoliation and Doping Process for Generating MX2 Nanosheets for Electrocatalytic Hydrogen Evolution Reaction, ACS Appl. Mater. Interfaces, 2019, 11(16), 14786–14795 CrossRef.
  98. X. Zhai, R. Zhang, H. Sheng, J. Wang, Y. Zhu, Z. Lu, Z. Li, X. Huang, H. Li and G. Lu, Direct Observation of the Light-Induced Exfoliation of Molybdenum Disulfide Sheets in Water Medium, ACS Nano, 2021, 15(3), 5661–5670 CrossRef.
  99. A. M. Jawaid, A. J. Ritter and R. A. Vaia, Mechanism for Redox Exfoliation of Layered Transition Metal Dichalcogenides, Chem. Mater., 2020, 32(15), 6550–6565 CrossRef.
  100. A. Jawaid, J. Che, L. F. Drummy, J. Bultman, A. Waite, M.-S. Hsiao and R. A. Vaia, Redox Exfoliation of Layered Transition Metal Dichalcogenides, ACS Nano, 2017, 11(1), 635–646 CrossRef CAS PubMed.
  101. D. Hu, C. Ye, X. Wang, X. Zhao, L. Kang, J. Liu, R. Duan, X. Cao, Y. He, J. Hu, S. Li, Q. Zeng, Y. Deng, P.-F. Yin, A. Ariando, Y. Huang, H. Zhang, X. R. Wang and Z. Liu, Chemical Vapor Deposition of Superconducting FeTe1−xSex Nanosheets, Nano Lett., 2021, 21(12), 5338–5344 CrossRef CAS.
  102. J.-K. Huang, N. Saito, Y. Cai, Y. Wan, C.-C. Cheng, M. Li, J. Shi, K. Tamada, V. C. Tung, S. Li and L.-J. Li, Steam-Assisted Chemical Vapor Deposition of Zeolitic Imidazolate Framework, ACS Mater. Lett., 2020, 2(5), 485–491 CrossRef CAS.
  103. X. Bai, Z. Xu, Q. Zhang, S. Li, Y. Dai, X. Cui, H. H. Yoon, L. Yao, H. Jiang, M. Du, Y. Zhang, E. I. Kauppinen and Z. Sun, Molybdenum Disulfide/Double-Wall Carbon Nanotube Mixed-Dimensional Heterostructures, Adv. Mater. Interfaces, 2022, 9, 2200193 CrossRef CAS.
  104. P. Wang, T. Zhao, S. Xiao, Q. Chen, Y. Zhang, L. Li, J. Wang, Y. Xue, B. Zhang, D. Li and B. Luo, Geometry-Derived Asymmetric Schottky Contacts Based on Chemical Vapor Deposited MoS2, ACS Appl. Electron. Mater., 2024, 6(10), 7475–7483 CrossRef CAS.
  105. A. Tarasov, P. M. Campbell, M.-Y. Tsai, Z. R. Hesabi, J. Feirer, S. Graham, W. J. Ready and E. M. Vogel, Highly Uniform Trilayer Molybdenum Disulfide for Wafer-Scale Device Fabrication, Adv. Funct. Mater., 2014, 24, 6389–6400 CrossRef CAS.
  106. D. Mouloua, K. Kaja, M. Lejeune, A. Zeinert, M. E. Marssi, M. A. E. Khakani and M. Jouiad, Scalable and Contactless Optical Dye Sensors Based on Differential Reflectivity of Excitonic Peaks by MoS2 Nanostructures, ACS Appl. Nano Mater., 2024, 7(8), 9366–9374 CrossRef CAS.
  107. H. Li, X. Duan, X. Wu, X. Zhuang, H. Zhou, Q. Zhang, X. Zhu, W. Hu, P. Ren, P. Guo, L. Ma, X. Fan, X. Wang, J. Xu, A. Pan and X. Duan, Growth of Alloy MoS2xSe2(1−x) Nanosheets with Fully Tunable Chemical Compositions and Optical Properties, J. Am. Chem. Soc., 2014, 136(10), 3756–3759 CrossRef CAS.
  108. H. Li, Q. Zhang, X. Duan, X. Wu, X. Fan, X. Zhu, X. Zhuang, W. Hu, H. Zhou, A. Pan and X. Duan, Lateral Growth of Composition Graded Atomic Layer MoS2(1−x)Se2x Nanosheets, J. Am. Chem. Soc., 2015, 137(16), 5284–5287 CrossRef CAS PubMed.
  109. Y. Huang, K. Yu, H. Li, K. Xu, Z. Liang, D. Walker, P. Ferreira, P. Fischer and D. (Emma) Fan, Scalable Fabrication of Molybdenum Disulfide Nanostructures and their Assembly, Adv. Mater., 2020, 32, 2003439 CrossRef CAS.
  110. L. Chen, M. Zhang, X. Yang, W. Li, J. Zheng, W. Gan and J. Xu, Sandwich-structured MnO2@N-doped Carbon@MnO2 nanotubes for high-performance supercapacitors, J. Alloys Compd., 2017, 695, 3339–3347 CrossRef CAS.
  111. V. Awasthi, S. Dey and P. Srivastava, Satish Kumar Dubey. α-MoO3 Microflakes Decorated with Gold Nanoislands as SERS-Active Substrates, ACS Appl. Nano Mater., 2024, 7(2), 2325–2334 CrossRef CAS.
  112. J. Dong, D. Ding, C. Jin, Y. Liu and F. Ding, Edge Reconstruction-Dependent Growth Kinetics of MoS2, ACS Nano, 2023, 17(1), 127–136 CrossRef CAS.
  113. T. Chowdhury, J. Kim, E. C. Sadler, C. Li, S. W. Lee, K. Jo, W. Xu, D. H. Gracias, N. V. Drichko, D. Jariwala, T. H. Brintlinger, T. Mueller, H.-G. Park and T. J. Kempa, Substrate-directed synthesis of MoS2 nanocrystals with tunable dimensionality and optical properties, Nat. Nanotechnol., 2020, 15, 29–34 CrossRef CAS.
  114. P. Yang, D. Wang, X. Zhao, W. Quan, Q. Jiang, X. Li, B. Tang, J. Hu, L. Zhu, S. Pan, Y. Shi, Y. Huan, F. Cui, S. Qiao, Q. Chen, Z. Liu, X. Zou and Y. Zhang, Epitaxial growth of inch-scale single-crystal transition metal dichalcogenides through the patching of unidirectionally orientated ribbons, Nat. Commun., 2022, 13, 3238 CrossRef CAS.
  115. K. C. Kwon, C. Kim, Q. V. Le, S. Gim, J.-M. Jeon, J. Y. Ham, J.-L. Lee, H. W. Jang and S. Y. Kim, Synthesis of Atomically Thin Transition Metal Disulfides for Charge Transport Layers in Optoelectronic Devices, ACS Nano, 2015, 9(4), 4146–4155 CrossRef CAS.
  116. A. M. v. d. Zande, P. Y. Huang, D. A. Chenet, T. C. Berkelbach, Y. You, G.-H. Lee, T. F. Heinz, D. R. Reichman, D. A. Muller and J. C. Hone, Grains and grain boundaries in highly crystalline monolayer molybdenum disulphide, Nat. Mater., 2013, 12, 554–561 CrossRef.
  117. J.-L. Wree, E. Ciftyurek, D. Zanders, N. Boysen, A. Kostka, D. Rogalla, M. Kasischke, A. Ostendorf, K. Schierbaum and A. Devi, A new metalorganic chemical vapor deposition process for MoS2 with a 1,4-diazabutadienyl stabilized molybdenum precursor and elemental sulfur, Dalton Trans., 2020, 49, 13462–13474 RSC.
  118. K. Kang, S. Xie, L. Huang, Y. Han, P. Y. Huang, K. F. Mak, C.-J. Kim, D. Muller and J. Park, High-mobility three-atom-thick semiconducting films with wafer-scale homogeneity, Nature, 2015, 520, 656–660 CrossRef CAS PubMed.
  119. C. Altavilla, M. Sarno and P. Ciambelli, A Novel Wet Chemistry Approach for the Synthesis of Hybrid 2D Free-Floating Single or Multilayer Nanosheets of MS2@oleylamine (M[double bond, length as m-dash]Mo, W), Chem. Mater., 2011, 23(17), 3879–3885 CrossRef CAS.
  120. R. Sanikop, A. K. Budumuru, S. Gautam, K. H. Chae and C. Sudakar, Robust Ferromagnetism in Li-Intercalated and -Deintercalated MoS2 Nanosheets: Implications for 2D Spintronics, ACS Appl. Nano Mater., 2020, 3(12), 11825–11837 CrossRef CAS.
  121. A. P. Frauendorf, A. Niebur, D. Rudolph, M. Oestreich, J. Lauth and J. Hübner, Ultrafast Recombination Dynamics under Lateral Confinement and Cryogenic Temperatures in Colloidal MoS2, J. Phys. Chem. C, 2024, 128(39), 16597–16606 CrossRef CAS.
  122. S. Lambora and A. Bhardwaj, Morphology Transition with Temperature and its Effect on Optical Properties of Colloidal MoS2 Nanostructures, ACS Omega, 2023, 8(30), 27725–27731 CrossRef CAS PubMed.
  123. S. Lambora and A. Bhardwaj, Enhanced Photoluminescence Quantum Yield of Colloidal MoS2 Quantum Dots by Optimization of Oleic Acid and Oleylamine for Light-Emitting Diode Applications, ACS Appl. Nano Mater., 2024, 7(1), 163–169 CrossRef CAS.
  124. Z. Liu, K. Nie, X. Qu, X. Li, B. Li, Y. Yuan, S. Chong, P. Liu, Y. Li, Z. Yin and W. Huang, General Bottom-Up Colloidal Synthesis of Nano-Monolayer Transition-Metal Dichalcogenides with High 1T′-Phase Purity, J. Am. Chem. Soc., 2022, 144(11), 4863–4873 CrossRef CAS PubMed.
  125. C. Meerbach, B. Klemmed, D. Spittel, C. Bauer, Y. J. Park, R. Hübner, H. Y. Jeong, D. Erb, H. S. Shin, V. Lesnyak and A. Eychmüller, General Colloidal Synthesis of Transition-Metal Disulfide Nanomaterials as Electrocatalysts for Hydrogen Evolution Reaction, ACS Appl. Mater. Interfaces, 2020, 12(11), 13148–13155 CrossRef CAS PubMed.
  126. J. J. Luo, L. Y. Qin, X. Y. Zan, H. L. Zou, H. Q. Luo, N. B. Li and B. L. Li, Cysteine-Induced Chirality Evolution of Molybdenum Disulfide Nanodots from a Bottom-Up Strategy, Langmuir, 2024, 40(29), 14900–14907 CrossRef CAS PubMed.
  127. J. Chen, G. Liu, Y.-z. Zhu, M. Su, P. Yin, X.-j. Wu, Q. Lu, C. Tan, M. Zhao, Z. Liu, W. Yang, H. Li, G.-H. Nam, L. Zhang, Z. Chen, X. Huang, P. M. Radjenovic, W. Huang, Z.-q. Tian, J.-f. Li and H. Zhang, Ag@MoS2 Core–Shell Heterostructure as SERS Platform to Reveal the Hydrogen Evolution Active Sites of Single-Layer MoS2, J. Am. Chem. Soc., 2020, 142(15), 7161–7167 CrossRef CAS PubMed.
  128. Q. Gong, L. Cheng, C. Liu, M. Zhang, Q. Feng, H. Ye, M. Zeng, L. Xie, Z. Liu and Y. Li, Ultrathin MoS2(1−x)Se2x Alloy Nanoflakes For Electrocatalytic Hydrogen Evolution Reaction, ACS Catal., 2015, 5(4), 2213–2219 CrossRef CAS.
  129. J. Li, A. Wrzesińska-Lashkova, M. Deconinck, M. Göbel, Y. Vaynzof, V. Lesnyak and A. Eychmüller, Facile and Scalable Colloidal Synthesis of Transition Metal Dichalcogenide Nanoparticles with High-Performance Hydrogen Production, ACS Appl. Mater. Interfaces, 2024, 16(28), 36315–36321 CrossRef CAS PubMed.
  130. E. J. Endres, J. R. B. Espano, A. Koziel, A. R. Peng, A. A. Shults and J. E. Macdonald, Controlling Phase in Colloidal Synthesis, ACS Nanosci. Au, 2024, 4(3), 158–175 CrossRef CAS PubMed.
  131. X. Geng, W. Sun, W. Wu, B. Chen, A. Al-Hilo, M. Benamara, H. Zhu, F. Watanabe, J. Cui and T.-p. Chen, Pure and stable metallic phase molybdenum disulfide nanosheets for hydrogen evolution reaction, Nat. Commun., 2016, 7, 10672 CrossRef CAS PubMed.
  132. N. Kumar, E. Fosso-Kankeu and S. S. Ray, Achieving Controllable MoS2 Nanostructures with Increased Interlayer Spacing for Efficient Removal of Pb(II) from Aquatic Systems, ACS Appl. Mater. Interfaces, 2019, 11(21), 19141–19155 CrossRef CAS PubMed.
  133. E. Saias, A. Ismach and I. Zucker, Engineering the Performance and Stability of Molybdenum Disulfide for Heavy Metal Removal, ACS Appl. Mater. Interfaces, 2023, 15(5), 6603–6611 CrossRef CAS.
  134. X. Li, S. Guo, J. Su, X. Ren and Z. Fang, Efficient Raman Enhancement in Molybdenum Disulfide by Tuning the Interlayer Spacing, ACS Appl. Mater. Interfaces, 2020, 12(25), 28474–28483 CrossRef CAS PubMed.
  135. A. Kaushik, J. Singh, R. Soni and J. Pratap Singh, MoS2−Ag Nanocomposite-Based SERS Substrates with an Ultralow Detection Limit, ACS Appl. Nano Mater., 2023, 6(11), 9236–9246 CrossRef CAS.
  136. H. Yu, Y. Chen, Z. Wen, R. Wang, S. Jia, W. Zhu, Y. Song, H. Sun and B. Liu, Selective SERS Sensing of R6G Molecules Using MoS2 Nanoflowers under Pressure, Langmuir, 2024, 40(41), 21804–21813 CrossRef CAS PubMed.
  137. H. Yu, H. Sun, J. Ma, B. Han, R. Wang, Y. Ma, G. Lou and Y. Song, Resonance-Assisted Surface-Enhanced Raman Spectroscopy Amplification on Hierarchical Rose-Shaped MoS2/Au Nanocomposites, Langmuir, 2024, 40(1), 380–388 CrossRef CAS PubMed.
  138. S. J. Panchu, K. Raju, P. Singh, D. D. Johnson and H. C. Swart, High Mass Loading of Flowerlike Ni-MoS2 Microspheres toward Efficient Intercalation Pseudocapacitive Electrodes, ACS Appl. Energy Mater., 2023, 6(4), 2187–2198 CrossRef CAS.
  139. H. Huang, C. Du, H. Shi, X. Feng, J. Li, Y. Tan and W. Song, Water-Soluble Monolayer Molybdenum Disulfide Quantum Dots with Upconversion Fluorescence, Part. Part. Syst. Charact., 2015, 32, 72–79 CrossRef CAS.
  140. Y. Gao, Z. Du, X. Jiang, J. Zheng, J. Jiang, S. Ye, B. Zou, T. Yang and J. Zhao, Broadband MoS2 Square Nanotube-Based Photodetectors, ACS Appl. Nano Mater., 2023, 6(8), 7044–7054 CrossRef CAS.
  141. H.-K. Lee, D. Sakemi, R. Selyanchyn, C.-G. Lee and S.-W. Lee, Titania Nanocoating on MnCO3 Microspheres via Liquid-Phase Deposition for Fabrication of Template-Assisted Core–Shell- and Hollow-Structured Composites, ACS Appl. Mater. Interfaces, 2014, 6(1), 57–64 CrossRef CAS PubMed.
  142. L. Yang, A. Zhang and L. Zhang, Light-Driven Fuel Cell with a 2D/3D Hierarchical CuS@MnS Z-Scheme Catalyst for H2O2 Generation, ACS Appl. Mater. Interfaces, 2023, 15(15), 18951–18961 CrossRef CAS PubMed.
  143. A. Taufik, Y. Asakura, T. Hasegawa and S. Yin, MoS2−xSex Nanoparticles for NO Detection at Room Temperature, ACS Appl. Nano Mater., 2021, 4(7), 6861–6871 CrossRef CAS.
  144. Y. Zhang, H. Tao, S. Du and X. Yang, Conversion of MoS2 to a Ternary MoS2−xSex Alloy for High-Performance Sodium-Ion Batteries, ACS Appl. Mater. Interfaces, 2019, 11(12), 11327–11337 CrossRef CAS PubMed.
  145. X. Zheng, H. Han, J. Liu, Y. Yang, L. Pan, S. Zhang, S. Meng and S. Chen, Sulfur Vacancy-Mediated Electron–Hole Separation at MoS2/CdS Heterojunctions for Boosting Photocatalytic N2 Reduction, ACS Appl. Energy Mater., 2022, 5(4), 4475–4485 CrossRef CAS.
  146. B. Chen, Q. Zhang, P. Zhao, M. Cen, Y. Song, W. Zhao, W. Peng, Y. Li, F. Zhang and X. Fan, Coupled Co-Doped MoS2 and CoS2 as the Dual-Active Site Catalyst for Chemoselective Hydrogenation, ACS Appl. Mater. Interfaces, 2023, 15(1), 1317–1325 CrossRef CAS PubMed.
  147. J. Dong, W. Fang, H. Yuan, W. Xia, X. Zeng and W. Shangguan, Few-Layered MoS2/ZnCdS/ZnS Heterostructures with an Enhanced Photocatalytic Hydrogen Evolution, ACS Appl. Energy Mater., 2022, 5(4), 4893–4902 CrossRef CAS.
  148. Y. Wang, J. Wang, Y. Zhao, Z. Zhao, R. Yu, S. Yang, Z. Guo, C. Tang and Y. Fang, Flower-Like CuS/MoS2Nanoflake Composites as Heterojunctions for Microwave Absorption, ACS Appl. Nano Mater., 2024, 7(7), 7237–7247 CrossRef CAS.
  149. L. Hu, Y. Ren, H. Yang and Q. Xu, Fabrication of 3D Hierarchical MoS2/Polyaniline and MoS2/C Architectures for Lithium-Ion Battery Applications, ACS Appl. Mater. Interfaces, 2014, 6(16), 14644–14652 CrossRef CAS PubMed.
  150. M. Zhu, Y. Yang and Y. Ma, Salt-assisted synthesis of advanced carbon-based materials for energy-related applications, Green Chem., 2023, 25, 10263–10303 RSC.
  151. Y. Yang, Y. Ma, S. Li and M. Zhu, Molten salt technique for the synthesis of carbon-based materials for supercapacitors, Green Chem., 2023, 25, 10209–10234 RSC.
  152. F. A. Cevallos, S. Guo, H. Heo, G. Scuri, Y. Zhou, J. Sung, T. Taniguchi, K. Watanabe, P. Kim, H. Park and R. J. Cava, Liquid Salt Transport Growth of Single Crystals of the Layered Dichalcogenides MoS2 and WS2, Cryst. Growth Des., 2019, 19(10), 5762–5767 CrossRef CAS.
  153. M. Suleman, S. Lee, M. Kim, V. H. Nguyen, M. Riaz, N. Nasir, S. Kumar, H. M. Park, J. Jung and Y. Seo, NaCl-Assisted Temperature-Dependent Controllable Growth of Large-Area MoS2 Crystals Using Confined-Space CVD, ACS Omega, 2022, 7(34), 30074–30086 CrossRef CAS PubMed.
  154. A. Singh, M. Sharma and R. Singh, NaCl-Assisted CVD Growth of Large-Area High-Quality Trilayer MoS2 and the Role of the Concentration Boundary Layer, Cryst. Growth Des., 2021, 21(9), 4940–4946 CrossRef CAS.
  155. M. Khan, R. Meena, D. K. Avasthi and A. Tripathi, Study of Ion Velocity Effect on the Band Gap of CVD-Grown Few-Layer MoS2, ACS Omega, 2023, 8(49), 46540–46547 CrossRef CAS PubMed.
  156. J. Lei, Y. Xie, A. Kutana, K. V. Bets and B. I. Yakobson, Salt-Assisted MoS2 Growth: Molecular Mechanisms from the First Principles, J. Am. Chem. Soc., 2022, 144(16), 7497–7503 CrossRef CAS PubMed.
  157. B. Li, K. Wang, S. He, H. Du, T. Wang, Z. Du and W. Ai, Boosted Hydrogen Evolution via Molten Salt Synthesis of Vacancy-Rich MoSxSe2−x Electrocatalysts, ACS Sustainable Chem. Eng., 2024, 12(12), 4867–4875 CrossRef CAS.
  158. Y. Jin, M. Cheng, H. Liu, M. Ouzounian, T. S. Hu, B. You, G. Shao, X. Liu, Y. Liu, H. Li, S. Li, J. Guan and S. Liu, Na2SO4-Regulated High-Quality Growth of Transition Metal Dichalcogenides by Controlling Diffusion, Chem. Mater., 2020, 32(13), 5616–5625 CrossRef CAS.
  159. I.-K. Oh, W.-H. Kim, L. Zeng, J. Singh, D. Bae, A. J. M. Mackus, J.-G. Song, S. Seo, B. Shong, H. Kim and S. F. Bent, Synthesis of a Hybrid Nanostructure of ZnO-Decorated MoS2 by Atomic Layer Deposition, ACS Nano, 2020, 14(2), 1757–1769 CrossRef CAS PubMed.
  160. D.-H. Zhao, X.-J. Guo, L. Tong, T.-Y. Wang, Z.-H. Gu, T.-B. Zhang, H. Zhu, W.-Z. Bao, L. Chen, L. Ji, Q.-Q. Sun and D. W. Zhang, Large-Scale Multilayer MoS2 Nanosheets Grown by Atomic Layer Deposition for Sensitive Photodetectors, ACS Appl. Nano Mater., 2022, 5(8), 10431–10440 CrossRef CAS.
  161. L. Cheng, X. Qin, A. T. Lucero, A. Azcatl, J. Huang, R. M. Wallace, K. Cho and J. Kim, Atomic Layer Deposition of a High-k Dielectric on MoS2 Using Trimethylaluminum and Ozone, ACS Appl. Mater. Interfaces, 2014, 6(15), 11834–11838 CrossRef CAS PubMed.
  162. M. S. M. Saifullah, M. Asbahi, M. B.-K. Kiyani, S. S. Liow, S. B. Dolmanan, A. M. Yong, E. A. H. Ong, A. I. Saifullah, H. R. Tan, N. Dwivedi, T. Dutta, R. Ganesan, S. Valiyaveettil, K. S. L. Chong and S. Tripathy, Room-Temperature Patterning of Nanoscale MoS2 under an Electron Beam, ACS Appl. Mater. Interfaces, 2020, 12(14), 16772–16781 CrossRef CAS PubMed.
  163. B. Dai, Y. Su, Y. Guo, C. Wu and Y. Xie, Recent Strategies for the Synthesis of Phase-Pure Ultrathin 1T/1T′ Transition Metal Dichalcogenide Nanosheets, Chem. Rev., 2024, 124(2), 420–454 CrossRef CAS PubMed.
  164. M. Kan, J. Y. Wang, X. W. Li, S. H. Zhang, Y. W. Li, Y. Kawazoe, Q. Sun and P. Jena, Structures and Phase Transition of a MoS2 Monolayer, J. Phys. Chem. C, 2014, 118(3), 1515–1522 CrossRef CAS.
  165. W. Zhao, X. Liu, X. Yang, C. Liu, X. Qian, T. Sun, W. Chang, J. Zhang and Z. Chen, Synthesis of Novel 1T/2H-MoS2 from MoO3 Nanowires with Enhanced Photocatalytic Performance, Nanomaterials, 2020, 10(6), 1124 CrossRef CAS PubMed.
  166. F. P. N. Antunes, V. S. Vaiss, S. R. Tavares, R. B. Capaz and A. A. Leitão, van der Waals interactions and the properties of graphite and 2H-, 3R- and 1T-MoS2: A comparative study, Comput. Mater. Sci., 2018, 152, 146–150 CrossRef.
  167. T. Wang, Q. Wang, W. Qu, Y. Wang, N. Hu, S. Wu, T. Feng, S. Song and F. Jia, The novel Auricularia-like MoS2 as binder-free electrodes for enhanced capacitive deionization, Mater. Res. Bull., 2024, 180, 113049 CrossRef CAS.
  168. T. Ying, Y. Xiong, H. Peng, R. Yang, L. Mei, Z. Zhang, W. Zheng, R. Yan, Y. Zhang, H. Hu, C. Ma, Y. Chen, X. Xu, J. Yang, D. Voiry, C. Y. Tang, J. Fan and Z. Zeng, Achieving Exceptional Volumetric Desalination Capacity Using Compact MoS2 Nanolaminates, Adv. Mater., 2024, 36, 2403385 CrossRef CAS PubMed.
  169. B. Pang, L. Xiang, K. Wang, S. Zeng, J. Guo and N. Li, Interlayer-expanded 1T-phase MoS2 as a cathode material for enhanced capacitive deionization, Desalination, 2025, 593, 118211 CrossRef CAS.
  170. J. Zhang, Y. Zhu, M. Tebyetekerwa, D. Li, D. Liu, W. Lei, L. Wang, Y. Zhang and Y. Lu, Vanadium-Doped Monolayer MoS2 with Tunable Optical Properties for Field-Effect Transistors, ACS Appl. Nano Mater., 2021, 4, 769–777 CrossRef CAS.
  171. H. Eidsvåg, P. Vajeeston and D. Velauthapillai, Doped MoS2 Polymorph for an Improved Hydrogen Evolution Reaction, ACS Omega, 2023, 8, 26263–26275 CrossRef PubMed.
  172. C. Lu, Y. Yang, S. Li and M. Zhu, Nanosheet floral clusters of Fe-doped Co3O4 for high-performance supercapacitors, Mater. Chem. Front., 2024, 8, 2282–2292 RSC.
  173. Z. Li, H. Zhou and G. Wang, Manganese doping strategy for enhanced capacitive deionization performance of MoS2, J. Environ. Chem. Eng., 2024, 12, 113277 CrossRef CAS.
  174. K. Sun, X. Yao, B. Yang, F. Jia and S. Song, Oxygen-incorporated molybdenum disulfide nanosheets as electrode for enhanced capacitive deionization, Desalination, 2020, 496, 114758 CrossRef CAS.
  175. W. Zhan, X. Zhang, Y. Yuan, Q. Weng, S. Song, M. d. J. Martínez-López, J. L. Arauz-Lara and F. Jia, Regulating Chemisorption and Electrosorption Activity for Efficient Uptake of Rare Earth Elements in Low Concentration on Oxygen-Doped Molybdenum Disulfide, ACS Nano, 2024, 18(9), 7298–7310 CrossRef CAS PubMed.
  176. F. Jia, K. Sun, B. Yang, X. Zhang, Q. Wang and S. Song, Defect-rich molybdenum disulfide as electrode for enhanced capacitive deionization from water, Desalination, 2018, 446, 21–30 CrossRef CAS.
  177. W. Zhan, Y. Yuan, X. Zhang, Y. Liang, S. Song, M. d. J. Martínez-López, J. L. Arauz-Lara and F. Jia, Efficient and selective Lead(II) removal within a wide concentration range through chemisorption and electrosorption coupling process via defective MoS2 electrode, Sep. Purif. Technol., 2024, 329, 125183 CrossRef CAS.
  178. X. Hu, Y. Li, Y. Xu, Z. Gan, X. Zou, J. Shi, X. Huang, Z. Li and Y. Li, Green one-step synthesis of carbon quantum dots from orange peel for fluorescent detection of Escherichia coli in milk, Food Chem., 2021, 339, 127775 CrossRef CAS PubMed.
  179. Z. Yang, T. Xu, H. Li, M. She, J. Chen, Z. Wang, S. Zhang and J. Li, Zero-Dimensional Carbon Nanomaterials for Fluorescent Sensing and Imaging, Chem. Rev., 2023, 123, 11047–11136 CrossRef CAS PubMed.
  180. M. Li, T. Chen, J. Gooding and J. Liu, Review of Carbon and Graphene Quantum Dots for Sensing, ACS Sens., 2019, 4(7), 1732–1748 CrossRef CAS PubMed.
  181. K. Zeng, W. Wei, L. Jiang, F. Zhu and D. Du, Use of Carbon Nanotubes as a Solid Support To Establish Quantitative (Centrifugation) and Qualitative (Filtration) Immunoassays To Detect Gentamicin Contamination in Commercial Milk, J. Agric. Food Chem., 2016, 64(41), 7874–7881 CrossRef CAS PubMed.
  182. A. V. Shkolin, I. E. Men'shchikov, E. V. Khozina, V. Y. Yakovlev, V. N. Simonov and A. A. Fomkin, Deformation of Microporous Carbon Adsorbent Sorbonorit-4 during Methane Adsorption, J. Chem. Eng. Data, 2022, 67(7), 1699–1714 CrossRef CAS.
  183. J. Fan, L. Duan, X. Zhang, Z. Li, P. Qiu, Y. He and P. Shang, Selective Adsorption and Recovery of Silver from Acidic Solution Using Biomass-Derived Sulfur-Doped Porous Carbon, ACS Appl. Mater. Interfaces, 2023, 15, 40088–40099 CrossRef CAS PubMed.
  184. S. Wang, J. Dou, T. Zhang, S. Li and X. Chen, Selective Adsorption of Methyl Orange and Methylene Blue by Porous Carbon Material Prepared From Potassium Citrate, ACS Omega, 2023, 8, 35024–35033 CrossRef CAS PubMed.
  185. H. Huang, Q. Liang, G. Li, H. Guo, Z. Wang, G. Yan, X. Li, H. Duan and J. Wang, Robust Spray Combustion Enabling Hierarchical Porous Carbon-Supported FeCoNi Alloy Catalyst for Zn–Air Batteries, ACS Appl. Mater. Interfaces, 2025, 17, 7763–7772 CrossRef CAS PubMed.
  186. A. Pandey, Twinkle, S. K. Sonkar and C. Dalal, Copper Nanoparticles-Decorated Carbon Nanoflakes as a Catalyst for Nitroaromatic Reduction, ACS Appl. Nano Mater., 2024, 7(22), 26238–26247 CrossRef CAS.
  187. K. Azam, S. Akhtar, Y. Y. Gong, M. N. Routledge, A. Ismail, C. A. F. Oliveira, S. Z. Iqbal and H. Ali, Evaluation of the impact of activated carbon-based filtration system on the concentration of aflatoxins and selected heavy metals in roasted coffee, Food Control, 2021, 121, 107583 CrossRef CAS.
  188. T. Di, Y. Xia, B. Pei, T. Zhu, T. Zhao, T. Li and L. Li, Preparation of Porous Carbon Materials Derived from Hyper-Cross-Linked Asphalt/Coal Tar and Their High Desulfurization Performance, Langmuir, 2020, 36, 11117–11124 CrossRef CAS PubMed.
  189. H. Mashhadimoslem, M. A. Abdol, K. Zanganeh, A. Shafeen, A. A. AlHammadi, M. Kamkar and A. Elkamel, Development of the CO2 Adsorption Model on Porous Adsorbent Materials Using Machine Learning Algorithms, ACS Appl. Energy Mater., 2024, 7(19), 8596–8609 CrossRef CAS.
  190. Z. Jing, J. Ding, T. Zhang, D. Yang, F. Qiu, Q. Chen and J. Xu, Flexible, versatility and superhydrophobic biomass carbon aerogels derived from corn bracts for efficient oil/water separation, Food Bioprod. Process., 2019, 115, 134–142 CrossRef CAS.
  191. T. Zhang, B. Zhao, Q. Chen, X. Peng, D. Yang and F. Qiu, Layered double hydroxide functionalized biomass carbon fiber for highly efficient and recyclable fluoride adsorption, Appl. Biol. Chem., 2019, 62, 12 CrossRef.
  192. T. Yin, Y. Zhang, D. Dong, T. Wang and J. Wang, Highly efficient capacitive removal of Cd2+ over MoS2-Carbon framework composite material in desulphurisation wastewater from coal-fired power plants, J. Cleaner Prod., 2022, 355, 131814 CrossRef CAS.
  193. Z. Hao, Y. Cai, Y. Wang, S. Xu and J. Wang, A coupling technology of capacitive deionization and MoS2/nitrogen-doped carbon spheres with abundant active sites for efficiently and selectively adsorbing low-concentration copper ions, J. Colloid Interface Sci., 2020, 564, 428–441 CrossRef CAS PubMed.
  194. T. Wu, X. Chen, H. Zhang, M. Zhao, L. Huang, J. Yan, M. Su and X. Liu, MoS2-encapsulated nitrogen-doped carbon bowls for highly efficient and selective removal of copper ions from wastewater, Sep. Purif. Technol., 2023, 304, 122284 CrossRef CAS.
  195. T. K. A. Nguyen, T.-H. Wang and R.-a. Doong, Architectures of flower-like MoS2 nanosheet coated N-doped carbon sphere electrode materials for enhanced capacitive deionization, Desalination, 2022, 540, 115979 CrossRef CAS.
  196. Z. Liu, B. Wei and L. Wang, Polyaniline-derived nitrogen-doped carbon/MoS2 nanocomposites as cathode for efficient hybrid capacitive deionization, Environ. Eng. Res., 2024, 29, 230204 CrossRef.
  197. H. Tian, J. Luo, J. Mu and B. Liu, Insight into the mechanisms of efficient selective removal of Pb2+ from wastewater by nano-flowered MoS2/NHCS electrodes, J. Environ. Chem. Eng., 2024, 12, 113039 CrossRef CAS.
  198. M. Zhu, J. Tang, W. Wei and S. Li, Recent progress in the syntheses and applications of multishelled hollow nanostructures, Mater. Chem. Front., 2020, 4, 1105–1149 RSC.
  199. C. Xu, D. Niu, N. Zheng, H. Yu, J. He and Y. Li, Facile Synthesis of Nitrogen-Doped Double-Shelled Hollow Mesoporous Carbon Nanospheres as High-Performance Anode Materials for Lithium Ion Batteries, ACS Sustainable Chem. Eng., 2018, 6(5), 5999–6007 CrossRef CAS.
  200. L. Ding, M. Zhang, Y. Ren, J. Xu, J. Zheng, H. Alsulami, M. A. Kutbi and F.-Y. Zhang, Carbon-Supported Nickel Nanoparticles on SiO2 Cores for Protein Adsorption and Nitroaromatics Reduction, ACS Appl. Nano Mater., 2020, 3(5), 4623–4634 CrossRef CAS.
  201. J. Hu, C. Shao, B. Li, Y. Yang and Y. Li, Fabrication of Carbonaceous Nanotubes and Mesoporous Nanofibers as Stable Anode Materials for Lithium-Ion Battery, ACS Appl. Nano Mater., 2018, 1(10), 5536–5542 CrossRef CAS.
  202. Y. Zhang, S. Fan, S. Gong, H. Liu, H. Xu, J. Qi, H. Wang, C. Li, W. Peng and J. Liu, Confinement and phase engineering boosting 1T phase MoS2/carbon hybrid for high-performance capacitive deionization, AIChE J., 2024, 70, e18417 CrossRef CAS.
  203. S. Tian, X. Zhang and Z. Zhang, Novel MoS2/NOMC electrodes with enhanced capacitive deionization performances, Chem. Eng. J., 2021, 409, 128200 CrossRef CAS.
  204. Q.-L. Zhu, W. Xia, L.-R. Zheng, R. Zou, Z. Liu and Q. Xu, Atomically Dispersed Fe/N-Doped Hierarchical Carbon Architectures Derived from a Metal–Organic Framework Composite for Extremely Efficient Electrocatalysis, ACS Energy Lett., 2017, 2, 504–511 CrossRef CAS.
  205. Z. Li, H. Zhou, W. Li, Z. Zou and G. Wang, Cooperative reinforcement of photocatalysis-coupled capacitive deionization in fructose intercalated MoS2 for removal of tetracycline, Sep. Purif. Technol., 2024, 331, 125583 CrossRef CAS.
  206. Z. Bin, L. Feng and Y. Yan, Biomimetic metalloporphyrin oxidase modified carbon nanotubes for highly sensitive and stable quantification of anti-oxidants tert-butylhydroquinone in plant oil, Food Chem., 2022, 388, 132898 CrossRef CAS PubMed.
  207. L. Meng, Y. Zeng and D. Zhu, Dynamic Liquid Membrane Electrochemical Modification of Carbon Nanotube Fiber for Electrochemical Microfabrication, ACS Appl. Mater. Interfaces, 2020, 12(5), 6183–6192 CrossRef CAS PubMed.
  208. Jyoti, T. Żołek, D. Maciejewska, E. Gilant, E. Gniazdowska, A. Kutner, K. R. Noworyta and W. Kutner, Polytyramine Film-Coated Single-Walled Carbon Nanotube Electrochemical Chemosensor with Molecularly Imprinted Polymer Nanoparticles for Duloxetine-Selective Determination in Human Plasma, ACS Sens., 2022, 7, 1829–1836 CrossRef CAS PubMed.
  209. X. Wang, Y. Xu, Y. Li, Y. Li, Z. Li, W. Zhang, X. Zou, J. Shi, X. Huang, C. Liu and W. Li, Rapid detection of cadmium ions in meat by a multi-walled carbon nanotubes enhanced metal-organic framework modified electrochemical sensor, Food Chem., 2021, 357, 129762 CrossRef CAS PubMed.
  210. P. Srimuk, J. Lee, S. Fleischmann, S. Choudhury, N. Jäckel, M. Zeiger, C. Kim, M. Aslan and V. Presser, Faradaic deionization of brackish and sea water via pseudocapacitive cation and anion intercalation into few-layered molybdenum disulfide, J. Mater. Chem. A, 2017, 5, 15640–15649 RSC.
  211. Y. Cai, W. Zhang, R. Fang, D. Zhao, Y. Wang and J. Wang, Well-dispersed few-layered MoS2 connected with robust 3D conductive architecture for rapid capacitive deionization process and its specific ion selectivity, Desalination, 2021, 520, 115325 CrossRef CAS.
  212. J. Han, T. Yan, J. Shen, L. Shi, J. Zhang and D. Zhang, Capacitive Deionization of Saline Water by Using MoS2–Graphene Hybrid Electrodes with High Volumetric Adsorption Capacity, Environ. Sci. Technol., 2019, 53(21), 12668–12676 CrossRef CAS PubMed.
  213. L. Gao, Q. Dong, S. Bai, S. Liang, C. Hu and J. Qiu, Graphene Oxide-Tuned MoS2 with an Expanded Interlayer for Efficient Hybrid Capacitive Deionization, ACS Sustainable Chem. Eng., 2020, 8(26), 9690–9697 CrossRef CAS.
  214. Y. Liu, J. Zhao, T. Bo, R. Tian, Y. Wang, S. Deng, H. Jiang, Y. Liu, G. Lisak, M. Chang, X. Li and S. Zhang, Enhanced Uranium Extraction via Charge Dynamics and Interfacial Polarization in MoS2/GO Heterojunction Electrodes, Small, 2024, 20, 2401374 Search PubMed.
  215. P. Zheng, L. Wang, Q. Wang and J. Zhang, Enhanced capacitive deionization by rGO@PEI/MoS2 nanocomposites with rich heterostructures, Sep. Purif. Technol., 2022, 295, 121156 Search PubMed.
  216. W. Peng, W. Wang, G. Han, Y. Huang and Y. Zhang, Fabrication of 3D flower-like MoS2/graphene composite as high-performance electrode for capacitive deionization, Desalination, 2020, 473, 114191 Search PubMed.
  217. W. Peng, W. Wang, M. Qi, Y. Miao, Y. Huang and F. Yu, Enhanced capacitive deionization of defect-containing MoS2/graphene composites through introducing appropriate MoS2 defect, Electrochim. Acta, 2021, 383, 138363 CrossRef CAS.
  218. Y. Liu, X. Du, Z. Wang, L. Zhang, Q. Chen, L. Wang, Z. Liu, X. Dou, H. Zhu and X. Yuan, MoS2 nanoflakes-coated electrospun carbon nanofibers for “rocking-chair” capacitive deionization, Desalination, 2021, 520, 115376 CrossRef CAS.
  219. Q. Wang, F. Jia, S. Song and Y. Li, Hydrophilic MoS2/polydopamine (PDA) nanocomposites as the electrode for enhanced capacitive deionization, Sep. Purif. Technol., 2020, 236, 116298 Search PubMed.
  220. J. Chen, Y. Sun, M. Zhang, J. Zhao, L. Song, L. Yan and G. Li, One-dimensional PPy microtubes/hybrid-phase MoS2 nanosheets for high-performance capacitive deionization, Desalination, 2024, 590, 117981 CrossRef CAS.
  221. M. Zhu, Q. Chen, J. Tang, W. Wei and S. Li, Core@shell β-FeOOH@polypyrolle derived N, S-codoped Fe3O4@N-doped porous carbon nanococoons for high performance supercapacitors, Appl. Surf. Sci., 2019, 480, 582–592 Search PubMed.
  222. H. Shehzad, J. Chen, M. T. Shuang, Z. Liu, Z. H. Farooqi, A. Sharif, L. Zhou, E. Ahmed, A. Irfan, R. Begum and J. Ouyang, Insights into electro-assisted and selective adsorption of U(VI) using hierarchical porous and activated biocarbon from lotus pods/2D-MoS2/polypyrrole composites through capacitive deionization, Process Saf. Environ. Prot., 2024, 181, 354–366 Search PubMed.
  223. H. Shehzad, M. T. Shuang, J. Chen, Z. Liu, A. Sharif, Z. H. Farooqi, E. Ahmed, R. Begum, L. Zhou, J. Ouyang, A. Irfan, A. R. Chaudhry, S. Shaukat and U. Hussain, Biomass-derived N, P-codoped templated biocarbon@2D-MoS2/polypyrrole based hybrids for U(VI) electrosorption, J. Environ. Chem. Eng., 2024, 12, 111957 CrossRef CAS.
  224. H. Shehzad, J. Chen, M. T. Shuang, Z. Liu, Z. H. Farooqi, A. Sharif, E. Ahmed, L. Zhou, A. Irfan, R. Begum, F. Iqbal and J. Ouyang, Fabrication of an efficient hierarchical mesoporous 2D-MoS2/CNT/polypyrrole based composite electrodes for competitive and selective U6+ removal using capacitive deionization: Mechanistic evaluation through cyclic voltammetry, Colloids Surf., A, 2024, 680, 132637 Search PubMed.
  225. R. Fang, L. Zhang, S. Huang, Y. Zhao, S. Zhang and Y. Wang, 3D heterostructure constructed by ion-doped polypyrrole coupled with MoS2@HCS and its highly efficient deionization performance, Sep. Purif. Technol., 2024, 339, 126606 CrossRef CAS.
  226. Z. Liu, B. Wei, K. Liu and L. Wang, DBS-doped polypyrrole/CNTs with 3D conductive architecture connected with MoS2 as symmetrical electrodes for boosted CDI capability, Sep. Purif. Technol., 2024, 345, 127380 Search PubMed.
  227. S. A. Zargar, M. D. M. Abadi, E. Soroush, A. M. Khachatourian, M. Golmohammad and A. Nemati, Synthesis of novel 2D/2D Ti3C2Tx MXene/1T-MoS2 heterostructure enhanced with carbon nanotubes as a highly-efficient electrode for hybrid capacitive deionization, J. Alloys Compd., 2024, 981, 173765 CrossRef CAS.
  228. S. Yao, D. Wang, W. Fu, X. Gao, S. Wang, Y. Liu, Z. Hou, J. Wang, K. Nie, J. Xie, Z. Yang and Y.-M. Yan, Unveiling the role of Atomic-Level “Pump-Driven” effect in MoS2/MnO2 for facilitating directional charge transfer in hybrid capacitive deionization, Chem. Eng. J., 2024, 490, 151608 Search PubMed.
  229. Z. Zhao, J. Zhao, Y. Sun, M. Ye and X. Wen, Synergetic pseudocapacitive sodium capture for efficient saline water desalination by iron oxide Hydroxide-Decorated palladium nanoparticle anchored 3D flowerlike molybdenum sulfide, Chem. Eng. J., 2023, 458, 141508 CrossRef CAS.
  230. J. Sun, Y. Li, H. Song, H. Li, Q. Lai, G. Egabaierdi, Q. Li, S. Zhang, H. He and A. Li, Enhanced capacitive deionization properties of activated carbon doped with carbon nanotube-bridged molybdenum disulfide, Chemosphere, 2023, 310, 136740 CrossRef CAS PubMed.
  231. J.-J. Huang, Y.-X. Zhang, W.-C. Huang, C.-S. Huang and D.-S. Wuu, MoS2/TiO2/graphite felt electrode preparation for capacitive deionization device performance improvement, Surf. Coat. Technol., 2024, 479, 130515 CrossRef CAS.
  232. M. Zhu, C. Lu and L. Liu, Progress and challenges of emerging MXene based materials for thermoelectric applications, Iscience, 2023, 26, 106718 CrossRef PubMed.
  233. M. Zhu and Y. Yang, An electrostatic self-assembly strategy to construct highly stable MXene/CuO nanocomposites for high performance supercapacitors, New J. Chem., 2025, 49, 5565–5569 RSC.
  234. J. Wu, B. Chen, A. Feng, K. Liu and Y. Yu, Electrostatically induced reconstruction of the 3D nitrogen-doped Ti3C2Tx electrode and its excellent desalination performance for hybrid capacitive deionization, New J. Chem., 2024, 48, 2496–2504 RSC.
  235. Y. Cai, Y. Wang, L. Zhang, R. Fang and J. Wang, 3D Heterostructure Constructed by Few-Layered MXenes with a MoS2 Layer as the Shielding Shell for Excellent Hybrid Capacitive Deionization and Enhanced Structural Stability, ACS Appl. Mater. Interfaces, 2022, 14(2), 2833–2847 CrossRef PubMed.
  236. Z. Chen, X. Xu, Y. Liu, J. Li, K. Wang, Z. Ding, F. Meng, T. Lu and L. Pan, Ultra-durable and highly-efficient hybrid capacitive deionization by MXene confined MoS2 heterostructure, Desalination, 2022, 528, 115616 CrossRef.
  237. H. Qie, M. Liu, X. Fu, X. Tan, Y. Zhang, M. Ren, Y. Zhang, Y. Pei, A. Lin, B. Xi and J. Cui, Interfacial Charge-Modulated Multifunctional MoS2/Ti3C2Tx Penetrating Electrode for High-Efficiency Freshwater Production, ACS Nano, 2022, 16(11), 18898–18909 CrossRef PubMed.
  238. S. Tian, X. Zhang and Z. Zhang, Capacitive deionization with MoS2/g-C3N4 electrodes, Desalination, 2020, 479, 114348 CrossRef.
  239. F. Zechel, P. Hutár, V. Vretenár, K Végsö, P. Šiffalovič and M. Sýkora, Green Colloidal Synthesis of MoS2 Nanoflakes, Inorg. Chem., 2023, 62(40), 16554–16563 CrossRef PubMed.

This journal is © the Partner Organisations 2025
Click here to see how this site uses Cookies. View our privacy policy here.