Next Article in Journal
Optimization of the Extraction of Antioxidant Compounds from Roselle Hibiscus Calyxes (Hibiscus sabdariffa), as a Source of Nutraceutical Beverages
Next Article in Special Issue
Preparation of Biochar with Developed Mesoporous Structure from Poplar Leaf Activated by KHCO3 and Its Efficient Adsorption of Oxytetracycline Hydrochloride
Previous Article in Journal
Influence of β-Cyclodextrin Methylation on Host-Guest Complex Stability: A Theoretical Study of Intra- and Intermolecular Interactions as Well as Host Dimer Formation
Previous Article in Special Issue
Ratiometric Near-Infrared Fluorescence Liposome Nanoprobe for H2S Detection In Vivo
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalytic Applications of ReS2-Based Heterostructures

Engineering Research Center of Ministry of Education for Geological Carbon Storage and Low Carbon Utilization of Resources, Beijing Key Laboratory of Materials Utilization of Nonmetallic Minerals and Solid Wastes, National Laboratory of Mineral Materials, School of Materials Science and Technology, China University of Geosciences (Beijing), Beijing 100083, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(6), 2627; https://doi.org/10.3390/molecules28062627
Submission received: 23 February 2023 / Revised: 10 March 2023 / Accepted: 10 March 2023 / Published: 14 March 2023
(This article belongs to the Special Issue Chemical Functionalization of Two-Dimensional Materials)

Abstract

:
ReS2-based heterostructures, which involve the coupling of a narrow band-gap semiconductor ReS2 with other wide band-gap semiconductors, have shown promising performance in energy conversion and environmental pollution protection in recent years. This review focuses on the preparation methods, encompassing hydrothermal, chemical vapor deposition, and exfoliation techniques, as well as achievements in correlated applications of ReS2-based heterostructures, including type-I, type-II heterostructures, and Z-scheme heterostructures for hydrogen evolution, reduction of CO2, and degradation of pollutants. We believe that this review provides an overview of the most recent advances to guide further research and development of ReS2-based heterostructures for photocatalysis.

Graphical Abstract

1. Introduction

Currently, the need for clean and renewable energy sources in order to mitigate the increasing environmental pollution and global energy crisis has become more urgent. Photocatalytic technology has emerged as a promising avenue for utilizing solar energy due to its low cost, abundance, and eco-friendly nature [1]. In 1972, Fujishima and Honda demonstrated for the first time the ability to produce hydrogen by splitting water on a TiO2 electrode [2]. Since then, TiO2 has been extensively studied due to its strong redox abilities, eco-friendliness, and chemical stability, which have made it a promising material for photocatalysis over the past few decades [3,4,5,6]. Despite its advantages, TiO2 suffers from a narrow spectral response range, a high carrier recombination rate, and a lack of catalytic active sites, thus limiting its photocatalytic hydrogen evolution (PHE) activity [7,8,9]. The emergence of two-dimensional (2D) photocatalysts, such as transition metal dichalcogenides (TMDs), has reinvigorated the progress in photocatalytic and photoelectrocatalytic research [10,11,12]. Due to their easy exfoliation into monolayers, TMDs provide numerous benefits such as a high surface area and decreased migration distance for photogenerated electron–hole pairs. This not only offers an increased number of reaction sites but also reduces the likelihood of electron–hole recombination, potentially improving photocatalytic performance [13]. Nevertheless, the electronic structure of TMDs is highly dependent on the number of layers [14]. Achieving precise control of the number of layers is critical to obtaining optimal catalytic performance, as exemplified by the fact that the photocatalytic hydrogen production of MoS2 is substantially improved when it converts from a multi-layer to a single-layer configuration [15]. As such, controlling the number of layers remains a daunting task that limits the utility of TMDs in photocatalytic applications [16,17].
As a newly emerged member of the 2D TMDs family, ReS2 has been widely studied in the fields of electronics [18,19], optoelectronics [20,21,22,23,24], and catalysis [25,26,27,28] due to its advantageous features such as a large surface area, tunable active sites, layer-independent electric/optical properties, and structural/vibrational anisotropy. ReS2 crystallizes in a twisted T phase with clusters of Re4 units, forming one-dimensional chains in each single layer [29]. Distinct from the conventional 2H phase TMDs materials (represented by MoS2 in Figure 1a) with high hexagonal symmetry, ReS2 belongs (Figure 1b) to the triclinic crystal system with only one symmetrical center, exhibiting a distorted 1T’ phase structure. The Re atoms show an octahedral coordination. The extra valence electron on the 5d orbital of Re results in a Peierls distortion, and the adjacent four Re atoms are linked by metallic bonds to form a Re4 cluster, which further distorts the original hexagon of ReS2 into a parallelogram [30]. It has been demonstrated that bulk ReS2 consists of electronically and dynamically decoupled layers bound by weak van der Waals forces, which gives rise to a pseudo single-layer architecture with nearly identical band structures and Raman active modes compared to that of its monolayer [29,31], consequently overcoming the difficulty of preparing large-area and single-crystal monolayers. Distinct from the indirect-to-direct band-gap transition of other group-VI TMDs, both single-layer ReS2 and multilayer ReS2 are shown to have a direct band gap of approximately 1.4 eV (Figure 1c), with suitable band-edge positions for photocatalytic water splitting [29,32]. Additionally, the weak interlayer coupling allows for easy preparation of nanosheets with ample surface area for co-catalysis with other materials. Lastly, ReS2 is significantly more adept at light absorption and carrier mobility than group-VI TMDs, thus making it an advantageous and effective photocatalyst under visible light illumination [32].
In 2018, Fu’s group undertook an experimental study on the optoelectronic properties of ReS2 and its performance in photocatalyzed hydrogen evolution [33]. The conduction and valence band edges of ReS2 are more closely spaced to the water redox potentials than CdS, thus rendering ReS2 more efficient for PHE. Moreover, the monolayer-like structure of the catalyst facilitates the formation of trions (a bound state of two electrons and one hole). These trions can then migrate to the surface and participate in two-electron reactions at the abundant active sites. Furthermore, the relatively intact basal planes of ReS2 can promote electron transport to the edge-reactive sites. Such a two-electron catalytic reaction endows ReS2 with a PHE rate of 13.023 mmol g−1 h−1 under visible light irradiation [34].
ReS2 is a versatile material but relatively less-explored among TMDs. It has garnered significant attention from the scientific community due to its unique properties compared to other TMDs. Rahman et al. critically analyzed the synthesis, structure growth mechanisms, properties, and contemporary applications of ReS2 in different areas [32]. Zhang et al. has highlighted the weak coupling characteristics of ReS2 [35], and the intrinsic anisotropic properties are summarized by Cao et al. [36]. Other recent reviews have summarized advances in ReS2 research with respect to electronics, optics, magnetism, and alkali-metal ion batteries [37,38,39,40]. Academic research regarding the photocatalytic properties of ReS2 materials is currently limited.
It is well established that heterostructures have numerous advantages in enhancing photocatalyst functions [1,41,42]. Rational alignment of ReS2 with other semiconductors into a tandem structure can increase the range of light absorption across the solar spectrum and optimize its photocatalytic performance by inhibiting charge carrier recombination as well as promoting charge separation [34,43,44,45,46]. In this review, we provide mainly an overview of the recent progress on ReS2 heterostructure, focusing on the synthesis methods and their photocatalytic applications. Firstly, the preparation methods are discussed, including hydrothermal reaction, chemical vapor deposition (CVD), and mechanical or chemical exfoliation. Furthermore, we focus mainly on photocatalytic applications based on ReS2-based heterostructure, such as photocatalytic hydrogen evolution, photocatalytic reduction of CO2, and pollutant degradation. Finally, the review is summarized, and the prospects are discussed from current views.

2. Synthetic Methods for ReS2 and Its Composite Material

Various synthesis techniques such as mechanical/chemical exfoliation, hydrothermal, and CVD methods have been developed to prepare ReS2-based photocatalysts with different morphologies including nanospheres, nanosheets, etc. [45,47], all of which can be used as core components in the construction of photocatalytic heterostructures.

2.1. Exfoliation from the Bulk ReS2

The weak interlayer coupling of ReS2 permits the separation of bulk ReS2 into nanosheets through exfoliation, which is a simple, high-yield method for ReS2 preparation [29,48,49]. Ultrathin nanostructures can be obtained by employing a hybrid of mechanical and liquid exfoliation. Bulk ReS2 produced using either commercial or hydrothermal synthesis can be dispersed in organic solvents (such as ethanol or N,N-dimethylformamide) through ultrasound-assisted liquid-phase exfoliation at 20–25 °C using an ice bath, followed by centrifugation and filtration to yield the nanosheets [34,47,50,51,52,53]. The bulk ReS2 can be exfoliated into a few layers (Figure 2a), and the original structure of ReS2 is maintained. Figure 2b,c show the intact basal plane and edge of ReS2 nanosheets, respectively [50]. Figure 2d,e display the ReS2 nanosheets with a thickness of about 4.20 nm [54]. The under-coordinated sulfur sites at the edges of ReS2 nanosheets serve as highly active centers for H2 generation, while the intact basal plane facilitates efficient electron transport to these edge S active sites for H2 evolution [34]. Chemical methods, such as the addition of diazonium salts, can also be employed to modify the surface structure of TMDs during the stripping process [55]. For example, the incorporation of hydroxyl functional groups such as carboxyl can enhance the hydrophilicity of ReS2 and improve its dispersibility in the ReS2−C6H5COOH/TiO2 physical mixture, thus preserving the photocatalytic active sites [56].

2.2. Hydrothermal Synthesis Reaction

The hydrothermal method is one of the simplest methods for the preparation of TMDs [16]. Specific precursors are used to carry out chemical reactions under high temperature and high pressure in a specially designed polytetrafluoroethylene-lined autoclave. For the synthesis of pristine ReS2, ammonium perrhenate (NH4ReO4), hydroxylamine hydrochloride (NH2OH·HCl), and thiourea (CS(NH2)2) are dissolved in deionized water. After stirring, they are transferred to a polytetrafluoroethylene autoclave for hydrothermal reaction by heating to 200–240 °C for 12–48 h [58,59,60]. After cooling, the ReS2 black powder is repeatedly washed with deionized water and ethanol, and the final product is collected after drying. The ReS2 prepared using this method typically displays the morphology of nanospheres, and the size of the ReS2 nanospheres can be adjusted by regulating the concentration of rhenium sources in the reaction [45].
The enhanced photocatalytic activity of ReS2 can be achieved by combining it with another semiconductor material to construct heterojunctions [61]. There are two ways to prepare ReS2-based heterojunctions via the hydrothermal method. The first is to incorporate the synthesized host semiconductor material into the precursor mixture for the preparation of ReS2, or vice versa, using a two-step hydrothermal method [46,62]. Wang et al. dispersed the generated TiO2 nanofibers into the ReS2 preparation solution for the hydrothermal reaction to anchor the ReS2 nanosheets on the porous TiO2 nanofibers to form TiO2@ReS2 composites. The few-layer ReS2 nanosheets with twisted 1T phase that were prepared via the hydrothermal method show large catalytic active sites, high electron mobility, and matching band structure [58]. The other is to fabricate a heterostructure with all the raw material in the one-pot hydrothermal method [63]. Chen et al. prepared a 1T’-ReS2/g-C3N4 heterostructure via the one-pot solvothermal reaction (Figure 2f). The obtained 1T’-ReS2 nanosheets can be converted to 2H-ReS2 after subsequent heat treatment (Figure 2g). The thin 1T’-ReS2 nanosheets with multi-layers grown on the g-C3N4 nanotubes (green circle in Figure 2h) show characteristic 2D-layered structures of TMDs with the single-layer thickness ~0.27 nm and the interlayer spacing ~0.65 nm (blue circle in Figure 2h). Alternatively, ReS2 prepared via the hydrothermal method can also be subsequently combined with other semiconductor materials through grinding or liquid-phase dispersion, thereby enabling the construction of heterojunctions [51,64].

2.3. Chemical Vapor Deposition Technique

ReS2 nanosheets prepared through hydrothermal and exfoliation methods contain numerous impurities that can act as electron-and-hole recombination centers, thus reducing the photocatalytic activity of ReS2 [65]. Compared to other methods, CVD-prepared TMDs have many advantages, including uniform appearance, controllable thickness, low defect rate, clean interface, and high synthesis efficiency [66,67]. On the one hand, the weak van der Waals interaction between TMDs makes it possible to stack them to form a variety of vertical heterostructures. On the other hand, chemical bonds can be formed at the interface to realize the interlocking of different 2D materials to form horizontal heterostructures with special interfacial properties. Vertical heterostructures can be prepared through methods such as mechanical exfoliation and transfer [68,69], direct CVD growth [70], electrochemical exfoliation [71], and liquid-phase exfoliation [47]. Horizontal heterostructures can generally be prepared only by the CVD approach [72,73,74].
The CVD preparation of ReS2 heterostructures can also be divided into a “one-step method” and a “two-step method”. The one-step method puts the precursor containing the elements of the synthesized heterostructure into the reaction system at one time. This method is based on the difference in the substrate adsorption energy or the control of growth conditions, such as growth temperature or carriers, to realize the growth of two materials one after another. For example, by utilizing the difference in adsorption energy between Re and W atoms on a Au substrate, a one-step synthesis of a vertical ReS2/WS2 heterojunction on a Au/W-Re alloy foil can be achieved [75]. The two-step method means that the growth of heterostructures is carried out in two separated steps. These methods can realize the high-quality and large-area preparation of single-crystal ReS2, which has been widely used in micro-nano electronic devices and photoelectric detection fields [22,23,76]. A substantial amount of work has reported the CVD preparation and growth mechanism of ReS2 heterostructure; readers can refer to the related papers [72,73,74,75,77,78] and reviews [79].
The majority of TMDs produced via CVD techniques exhibit a parallel growth relative to the underlying substrate, resulting in a flat film morphology. Under similar CVD growth conditions, ReS2 nanosheets display a propensity to bend and adopt a vertically oriented configuration on the underlying substrate [80]. Specifically, the vertical growth of two-dimensional ReS2 is highly spontaneous and independent of the substrate [81]. The high packing density (∼108 cm−2) of the ReS2 nanosheets promotes a substantial number of active Re atoms to be exposed at the sheet edges, providing abundant active sites for catalytic processes [33,70,80,82,83,84]. The common precursors for the CVD synthesis of ReS2 are ammonium rhenate (NH4ReO4), Re powder [82], and rhenium-containing oxides (ReO3 or Re2O7), with sulfur powder used as the predominant sulfur source and heated in a different zone [33,85,86]. The ReS2 prepared via the CVD method exists mostly in petal-like ReS2 nanowalls arranged vertically without agglomeration [28,33,87]. Zhang et al. anchored the ReS2 nanowalls vertically on the surface of the Si/SiO2 substrate (Figure 2i), which can distribute ultra-uniformly with a dense structure over a wide area (Figure 2j) [33]. The average length of the ReS2 nanowalls can be up to 200 nm with domain sizes of about 10 layers. The interplanar distance of the ReS2 nanowalls is about 0.62 nm (Figure 2k), corresponding to the (001) crystal plane. The beneficial exposure of the (001) facet ensured that more active sites were accessible for hydrogen adsorption, consequently enhancing the catalytic activity. The growth of the ReS2 single-crystal nanosheets does not need a polished substrate or a small lattice mismatch [87]. Cheng et al. grew ReS2 tightly on the TiO2-coated b-Si substrate with the shape of nanopyramids via the CVD method. The obtained ReS2 nanosheet was of high quality and exhibited a good crystal structure without defects [28]. Before the characterization of photocatalytic properties, ReS2 generally needs to be scraped off from the substrate, and then via the solution reaction (such as hydrothermal or self-assembly) to prepare heterostructures.

3. Heterostructured ReS2 Composites

The heterojunctions formed between the host semiconductors generate an internal electric field that facilitates the separation of electron–hole pairs and accelerates the migration of carriers, which is constructive for improving the photocatalytic efficiency [88,89,90]. ReS2-based heterostructures can fulfill the aforementioned requirements and demonstrate outstanding photocatalytic performance. According to the bandgaps and the electronic affinity of semiconductors, ReS2-based heterostructures can be divided into three different types: type-I (straddling gap), type-II (staggered gap), and Z-scheme system, the band alignment of which is shown in Figure 3. The band gap, the electron affinity, and the work function of the combined semiconductors can determine the dynamics of the electron and hole in the semiconductor heterojunctions.

3.1. Type I Heterostructure

In a type-I heterostructure (shown in Figure 3a), the photo-excited electrons and holes in the wide-gap material transfer to the narrow-gap material (ReS2), as demonstrated by the arrows [46,50,51,53,54,91,92,93]. The quantum confinement of the accumulated electrons and holes in the same semiconductor facilitates high carrier recombination rate and radiative recombination, which is incredibly detrimental to photocatalytic activity but desirable in light-emitting applications [91,94]. To inhibit recombination of photo-generated charge, a cocatalyst such as Pt that is deposited in situ onto the surface [54] or abundant sulfur vacancies [93] can act as an electron trapper to facilitate electron accumulation and transfer. Additionally, the abundant edge reaction sites in ultrathin ReS2 nanosheets in the composites can also accumulate electrons, thus accelerating reaction kinetics of the hydrogen evolution [53].

3.2. Type II Heterostructure

In a type-II heterojunction semiconductor, both valence band (VB) and conduction band (CB) of semiconductor 1 are situated at higher energy levels than those of semiconductor 2 (Figure 3b). The alignment of the energy levels facilitates electrons transferring from semiconductor 1 to semiconductor 2, while the holes migrate in the opposite direction. If both semiconductors are in close contact, an effective charge separation will take place upon light illumination. This leads to a decrease in charge recombination and an increase in the lifetime of charge carriers, resulting in higher photocatalyst activity. As one of the typical systems, the VB and CB of MX2 are both higher than that of ReS2. They can form a Type-II heterojunction structure in which the electrons transfer to ReS2 with holes transferred to MX2 [21,95,96]. Zhang et al. constructed a type-II heterojunction between ReS2 and XS2 (X = Mo, W), which can promote the electron transfer to further improve the photocatalytic performance of the counter electrode in dye-sensitized solar cells [96].
Traditional type-II heterojunctions (including n–n junctions and p–n junctions) have several obvious limitations. (1) From the thermodynamic point of view, the improvement of charge separation efficiency is at the expense of oxidation/reduction ability. Specifically, photo-generated electrons will accumulate on the CB of semiconductor 2 with a weaker reduction potential, and photo-generated holes will accumulate on the VB of semiconductor 1 with a lower oxidation potential. Following the type-II charge transfer mechanism, it will inevitably weaken the thermodynamic driving force of the photocatalytic reaction. (2) In terms of dynamics, the transition electrons of semiconductor 2 and the electrons transferred from semiconductor 1 will produce repulsive force. Similarly, the holes between them will also repel each other, which greatly hinders the continuous charge transfer in the type-II heterostructure. (3) Finally, the electrostatic attraction of electrons in semiconductor 2 and holes in semiconductor 1 also render the interfacial charge transfer of type-II composites difficult to achieve [61].

3.3. Z-Scheme Heterojunction

Thanks to inspiration from natural photosynthesis, in which two-step photoexcitation and electron transfer are connected in series to form a letter “Z”, similar artificial bionic Z-type heterostructures have been studied and developed in the photocatalysis field [97,98]. The first generation of the liquid-phase Z-scheme system has shuttle oxidation/reduction medium pairs, which are played by electron acceptor/donors, respectively, while there is no physical contact between oxidized semiconductors and reduced semiconductors. However, the first-generation Z-scheme photocatalytic system is limited to use in liquid-phase reactions, severely constraining its practical applications in gas- or solid-phase processes [99]. The second-generation all-solid-state Z-scheme heterojunction uses solid conductors (generally noble metal nanoparticles) instead of redox pairs to make it suitable for both liquid–solid and gas–solid reactions (Figure 3c). After the two semiconductors are excited at the same time, the electrons on the CB of the oxidized semiconductor will migrate to the solid conductor and then transfer to the VB of the reduced semiconductor. The length of this charge transfer path is significantly shortened compared to the initial liquid-phase Z-type heterostructure, which can greatly accelerate the charge transfer and separation efficiency. However, due to the difficulty in accurately controlling the heterojunction structure, the difficulty in directional electron transfer on the interface, and the competition between the light absorption of the mediator itself and the main catalytic components, these two heterostructures have not achieved the ideal photocatalytic effect [100].
The third-generation direct Z-type heterojunction is composed of only a reduced semiconductor and an oxidized semiconductor with a matching band structure and an ohmic contact at the interface (Figure 3d). This ohmic contact interface has certain defects and can be used as a recombination center for the electrons on the CB of the oxidized semiconductor and the holes on the VB of the reduced semiconductor. In addition, the impurity orbits contribute quasi-continuous energy levels and become charge transfer mediators. In addition to inheriting the advantages of the first two generations of Z-scheme systems, the direct Z-scheme without a solid medium greatly reduces the construction cost. Meanwhile, the direct Z-type heterostructure overcomes the light-shielding effect caused by redox pairs or noble metal particles [101]. Although the resistance of the direct Z-scheme at the solid–solid interface is generally considered to be higher than that of the all-solid-state Z-scheme, it is often possible to improve the conductivity of the interface by regulating or optimizing the chemical-binding force through molecular polarization to maximize the advantages of the Z-scheme heterojunction [56].

4. Photocatalytic Applications of Heterostructure ReS2

The direct band gap of ReS2 in both monolayer and bulk form is approximately 1.4 eV [29], allowing for the absorption of light across the entire visible spectrum and into the near-infrared region, making it an ideal material for visible light photocatalysis. The weak interlayer coupling of ReS2 results in optoelectronic properties that are largely independent of the number of layers, and the unsaturated edge sites are particularly exposed [35]. These advantages endow ReS2 with considerable potential as an efficient photocatalyst or cocatalyst. Considering the limitations of a single photocatalyst, coupling other semiconductors with ReS2 into heterojunctions can not only improve the utilization of solar energy but also reduce electron–hole pair recombination, thereby improving redox capacity. ReS2-based heterostructure composites have provoked extensive research attention and exhibit excellent photocatalytic activity in the fields of hydrogen production, CO2 reduction, and pollutant degradation.

4.1. Photocatalytic Hydrogen Production

Hydrogen, which is generated from photocatalytic water splitting, has been regarded as a potential source of energy for the development of a sustainable economy and society. Qualified photocatalysts should have a bandgap over 1.23 eV. The conduction band minimum (CBM) should be higher in energy than the H+/H2 water reduction potential, while the valence band maximum (VBM) should be lower in energy than the OH/O2 water oxidation potential. The band-gap and band-edge positions of monolayer and multilayer ReS2 are highly compatible with the energy levels required for water splitting [102]. Pioneer work by Zhang et al. reported that a multilayer structure ReS2 with monolayer-like features can generate a tremendous number of trions consisting of two electrons and one hole. This can trigger a two-electron catalytic reaction, offering a PHE rate of 13 mmol g−1 h−1 under visible light, which is larger than that of most reported TMD composite photocatalysts [33]. Due to its distorted 1T phase structure, transition metal σ bond, and direct band gap, ReS2 can serve as an electron reservoir with fast electron transport speed and active electronic states. Consequently, ReS2 has been widely recognized as an excellent cocatalyst for constructing a heterostructure with classical photocatalysts such as TiO2 [27,34,58,103,104], CdS [44,45,59,62], and g-C3N4 [46,54,57,92]. It can not only markedly enhance the visible light absorption, charge transfer, and carrier separation efficiency but also offer abundant edge active sites. Additionally, when Re vacancy is produced in ReS2, some unsaturated S atoms around the defect sites can promote the adsorption of a H+ proton, thus improving the HER performance more than perfect ReS2 [34,53,60]. The performance for ReS2 as a cocatalyst has been reported to be comparable to widely used noble-metal cocatalysts such as Pt [44,57]. Various photocatalysts have been summarized in Table 1 on the water splitting for the H2 production.
The morphology and structure of cocatalysts are important factors that can influence the photocatalytic performance. Cocatalysts typically exhibit morphologies such as nanoparticles, nanospheres, flakes, clusters, nanotubes, and quantum dots. Inspired by nature, Lin et al. designed novel sea-urchin-like ReS2 nanosheet/TiO2 nanoparticle heterojunctions through a simple one-pot hydrothermal method (Figure 4). Compared with the bulk ReS2 cocatalyst composed of a large number of agglomerated nanoparticles or nanosheets, the sea-urchin-like ReS2 cocatalyst shows a significantly accelerated charge transfer due to the unusual charge edge-collection effect. The photocatalytic H2 evolution rate is 3.71 mmol h−1 g−1 (an apparent quantum efficiency (AQE) of 16.09%), which is 231.9 times that of P25 TiO2 [104].
TMDs with a metallic 1T phase and semi-metallic 1T’ phase generally have high electrical conductivity [106,107]. However, most TMDs with the 1T-phase and 1T’-phase are thermodynamically unstable and can easily convert to the sTable 2H-phase [108,109]. Inspiringly, ReS2 has a thermodynamically sTable 1T’-phase crystal structure, which shows great promise for better catalytic performance [32,58]. As cocatalysts, the conductive 1T’-ReS2 nanosheets can trigger the host g-C3N4 nanotubes to generate more photoexcited charges, promote the carrier migration and separation, and supply abundant active sites for photocatalysis. Thus, the 1T’-ReS2/g-C3N4 composite containing 12 wt% 1T’-ReS2 exhibits an excellent photocatalytic H2 evolution rate of 2275 mmol h−1 g−1 [57]. Regulating phase structure and constructing an in-plane heterostructure of ReS2 can also provide more active sites and increase electron transfer [26]. Chen et al. reported that simply phosphorization treatment led to a shift in the Fermi level due to the extra electrons filling in the d orbital, and thus, ReS2 partially transformed from the T’ to the T phase [62]. The developed 2% ReS2-CdS/P-0.2 photocatalyst with in-plane T-T’ heterophase junctions enhanced the interfacial effect between CdS and ReS2, which can facilitate the electron transfer and exhibit an optimal hydrogen generation rate of 14.68 mmol g−1 h−1.
Heterostructures with well-identified interfaces are crucial for enhancing photocatalytic performance. The intimate interfacial contact between the individual components via the covalent bond can greatly accelerate the separation of photogenerated electrons and holes. Taking the TiO2@ReS2 nanocomposites for example, the photoexcited electrons can rapidly transfer to ReS2 and diffuse to the active sites due to the chemical interaction of the Ti−O−Re bond [58]. The optimized TiO2@ReS2 nanocomposites with 30 wt% ReS2 exhibit superior photocatalytic H2 evolution activity with a rate of 1404 μmol h−1 g−1. In another 2D/2D heterostructure, the cocatalyst ReS2 with abundant active sites is strongly anchored on the TiO2 by the chemical interaction of the Ti–S bond, and the established intimate interfacial contact facilitates efficient carrier transfer across the interface [103]. Recently, the strongly coupled interface between atomic-level regulated ReSe2 and various semiconductor photocatalysts and the abundant Re/Se active sites have been demonstrated to boost the performance for photocatalytic H2 evolution [110].
Chemical modification can adjust the optical and electronic properties of ReS2 nanosheets and produce new catalytic properties by increasing the catalytic active sites exposed in solution, which is an important goal for the development of practical and efficient solar hydrogen production devices. Wu et al. used C6H5COOH on ReS2 to form a molecular linkage between 2D ReS2 and 0D two-phase TiO2 through carboxylate (Figure 5) [56]. This structure provides a Z-scheme molecular electron channel, while the non-functionalized ReS2 provides an unstable and inefficient type-II path. The ReS2-C6H5COOH-TiO2 composite photocatalyst can not only provide good wettability, connection, and charge transport for solar hydrogen production but also prevent ReS2 from being oxidized by photogenerated holes. The composite photocatalyst exhibited high photocatalytic activity for solar hydrogen production (9.5 mmol h−1 g−1 ReS2 nanosheets, 4750 times higher than bulk ReS2) and high cycle stability over 20 h.

4.2. Photocatalytic CO2 Reduction

Anthropogenic emissions of greenhouse gases, such as CO2, into the environment can lead to global warming. To reduce CO2 production, various strategies have been developed. Using clean renewable energy to fix CO2 can promote the carbon cycle and alleviate the global energy crisis. Therefore, photocatalysis has received extensive attention because it can convert carbon dioxide into renewable fuels. Various photocatalysts have been summarized in Table 2 for the CO2 reduction.
The valence band edge of ReS2 (Ev = 1.36 V vs. NHE) is more positive than the oxidation potential of H2O oxidation (0.82 V vs. NHE at pH = 7), while its conduction band edge (EC = −0.20 V vs. NHE) is much lower than the reduction potential of CO2 (−0.52 V vs. NHE at pH = 7), indicating that ReS2 on its own does not have the ability of photocatalytic reduction of CO2 [111]. However, the introduction of defects on the ReS2 surface can not only modulate its band structure but also activate the inert surface to realize efficient charge transfer [51]. For example, Zhang et al. synthesized a type I heterojunction which was composed of ReS2 nanosheets and CdS nanoparticles via a self-assembly approach (Figure 6a–c). ReS2 with S vacancies induced by visible-light illumination is beneficial to the chemical adsorption and activation of CO2, as well as the electron transfer between CO2 and ReS2. The optimized ReS2/CdS heterostructure exhibits an enhanced photocatalytic CO2-to-CO conversion activity of 7.1 μmol g−1 accompanied by a prominent selectivity of 93.4% [51].
Owing to the excellent light absorption, large specific surface area, and abundant active sites of ReS2, it can serve as an ideal cocatalyst for photocatalytic CO2 reduction [51,64,111]. Huang et al. constructed an indirect Z-scheme heterostructure Au-Pt/Cu2O/ReS2 via chemical reduction and photodeposition (Figure 6d,e). The photocatalytic efficiency of the prepared heterojunction was significantly improved, and the CO yield was 30.11 μmol g−1, which was higher than that of pure Cu2O quantum dots (22.87 μmol g−1). In addition, the selectivity of CH4 and CO can be controlled by manipulating the mass ratio of Au/Pt (Figure 6f,g). Because of the high work function and excellent conductivity of Au and Pt, the electron transfer was accelerated between Cu2O and ReS2 (Figure 6h) [111].

4.3. Degradation of Organic Pollutants

The increase in energy consumption associated with the rapid development of modern society has led to a corresponding rise in environmental pollution, particularly water pollution, which poses a serious threat to human health and global ecological balance. To address this issue, advanced energy production and pollution control technologies should be developed. Among them, photocatalysts have demonstrated their efficacy in degrading organic pollutants. Various photocatalysts have been summarized in Table 3. Specifically, due to the nonbonding d electron of the Re atom, the 1T-ReS2 phase exhibits a diamond-chain structural distortion caused by the Re−Re bond interaction [112]. Therefore, the transitions of exciton in ReS2 are highly polarization-dependent [113]. More electron–hole pairs and a longer carrier lifetime parallel to the diamond-chain direction give rise to a faster photodegradation rate constant up to 12 times greater than that perpendicular to the diamond-chain direction [114].
The construction of a p-n heterojunction can induce a charge diffusion near the interface in nanocomposites until equilibrium is reached at the Fermi level between the two materials. For example, in the TiO2@ReS2 nanocomposites, this effect will lift the conduction band position of ReS2 higher than that of TiO2, leading to a better electron−hole separation efficiency. Under sunlight irradiation, the degradation activity of organic pigments can be significantly improved compared with pure TiO2 nanoparticles [47]. Similar strategies take effect such that the ReS2/MIL-88B(Fe) heterojunction can extend the lifetime of photoexcited electron−hole pairs, absorb more visible light, and enhance persulfate activation, leading to a better catalytic capacity for ibuprofen degradation compared with the individual component [52]. Electrostatic attraction is significant for fabricating heterojunctions. For instance, the g-C3N4 and defect-rich ReS2 attain close contact, resulting in faster generation of the reactive oxygen species than pristine g-C3N4 [116].
It is an efficient strategy to improve the photocatalytic performance by constructing a dynamic built-in electric field in a “self-driven” manner to continuously drive the separation of carriers [117]. Recently, a BaTiO3@ReS2 piezo-assisted photocatalysis heterostructure has been successfully synthesized by Liu et al. [115] through the coupling of interfacial covalent bonds and piezotronics, showing high degradation efficiency for refractory pollutants. The introduction of distorted 1T-ReS2 can enhance the light absorption of the hybrid nanostructure. Moreover, the formed interfacial Re-O covalent bond (Figure 7a,b) can serve as the channels for charge carriers’ transfer. In addition, the internal polarization field in the Schottky interface can reduce the Schottky barrier of photogenerated charge transfer. Consequently, the BaTiO3@ReS2 system obtained high O2 activation activity and an excellent synergistic effect under ultrasonic and light conditions. The BaTiO3@ReS2 exhibits ultrahigh pollutant degradation activity. The apparent rate constant of piezoelectric-assisted photocatalysis is 0.133 min−1, which is 16.6 and 2.44 times that of piezoelectric catalysis and photocatalysis, respectively. (Figure 7c,d) [115].

4.4. Photocatalytic Reduction of Metal Ions

Heavy metals in aquatic systems and drinking water sources are acutely toxic and carcinogenic to most organisms. Taking Cr (VI) as an example, photocatalytic reduction of Cr (VI) to low-toxic Cr (III) is generally considered to be a promising and practical option because this method is more effective and low-cost and does not produce any hazardous chemicals. Zhou et al. proposed a simple and feasible atomic hybridization strategy to accelerate the reaction kinetics process via constructing electronic pathways [86]. The reduced carbon quantum dots (rCQDs)/ReS2 heterostructure is synthesized via the CVD and hydrothermal method (Figure 8a,b). Compared with pure ReS2 nanosheets and CQDs/ReS2, the reduction reaction rate constants of the pseudo-first-order kinetic model were increased by about 13.1 and 4.3 times, respectively. (Figure 8c). C=O double bonds reduced and then anchored onto ReS2 nanosheets can form an electronic channel via Re-5d and O-2p orbital hybridization (Figure 8d), in which a photoinduced carrier of the rCQDs can freely migrate to ReS2 for hexavalent chromium reduction (Figure 8e) [86].

4.5. Photocatalytic Water Disinfection

The treatment cost of freshwater pollution is high, and the shortage of freshwater resources has become a more complex problem. Photocatalysis techniques have been regarded as one of the most promising strategies for environmental remediation. It is well known that when the photocatalyst absorbs light, electron−hole pairs are formed, which interact with water and dissolved oxygen to generate activated oxygen species (ROS), such as hydroxyl radicals, singlet oxygen, and superoxide [118]. These strong oxidants can destroy essential macromolecules in bacteria, resulting in bacterial disinfection [119]. The field of ReS2 photocatalytic sterilization is still in the preliminary stages. Pioneer work by Ghoshal et al. proved that vertically oriented ReS2 nanosheets show great potential for photocatalytic water disinfection [78]. However, due to the rapid recombination of the carriers, its inactivation efficiency still needs to be further enhanced.
It has been found that the construction of a 0D/2D heterostructure can greatly improve the sterilization efficiency of a photocatalyst because of the effective separation of charge carriers. Recently, Zhang et al. have reported that a Ag-modified ReS2 heterostructure grown on a series of substrates can be used as a visible-light-driven photocatalyst for highly efficient water sterilization. Different substrates and the size of Ag nanoparticles are critical to the photocatalytic bactericidal performance (Figure 9a,b). The Ag nanoparticles on the surface of ReS2 nanosheets can not only act as electron capture agents to promote the separation of photogenerated carriers but also inhibit the recombination of photogenerated carriers (Figure 9c). The main active substances in the photocatalysis process are h+, •OH, and •O2−. All Escherichia coli can be inactivated within 30 min at a concentration of 1.72 mg L−1 by the reusable Ag/ReS2 heterostructure under visible light irradiation (Figure 9d) [85].

5. Summary and Outlook

Two-dimensional ReS2, a special new material in the family of TMDs, has great potential as a photocatalyst due to its favorable band structure, which is independent of the number of layers, as well as its high specific surface area and numerous active sites. This review summarizes the basic properties and heterostructure synthesis methods of ReS2. Several types of ReS2-based heterostructures and their recent progress in photocatalysis fields, such as photocatalytic hydrogen production, CO2 reduction, and pollutant degradation, are then reviewed, demonstrating that ReS2 is an efficient and stable photocatalyst. However, the research on ReS2 is still in its early stages, and there is room for improvement of its photocatalytic performance. Based on synthetic chemical mechanisms, semiconductor physics of photocatalysis, and recent research results, the following aspects can be further explored for the rational design of efficient heterostructure photocatalysts:
  • Developing new synthetic methodologies to achieve precise synthesis of ReS2 and its composites. Mechanical mixing is usually used in the construction of heterojunctions for ReS2, but precise design for the interactions between the two or more materials has not yet been realized in the preparation of photocatalysts, resulting in random morphology and exposed crystal faces of the composites. With the growth mechanism of ReS2 as a basis, rational control of the growth conditions is expected to precisely control the size and crystal face orientation. In addition, reducing the size of ReS2 down to sub-10-nanometers is another intriguing direction due to the enhanced quantum confinement effect and derived novel photonic properties; however, structural controllability including particle diameter, edge structure, phase transition, etc. is highly important and is also well deserving of more research attention.
  • Surface engineering strategies toward tailorable physicochemical properties of ReS2. Few studies on the design and modification of the surface structure of ReS2 have been reported for photocatalytic research. Surface defects can become centers of the electron−hole complex, but they are not entirely inutile. For example, introducing Re vacancies in a reasonable way can enhance the adsorption ability of H+. In addition, organic molecules can form chemical bonds on the surface S vacancies, which can improve the interface wettability of ReS2 on the one hand and fine-tune its energy band position on the other hand.
  • Combining novel characterization methods with theoretical calculations. The surface and interface structure changes in photocatalysis should be thoroughly investigated from both microscopic and transient aspects. Real-time monitoring of intermediates and catalytic products is essential for understanding the photocatalytic mechanism and further optimizing the performance of photocatalysts. For instance, in situ Fourier Transform Infrared spectroscopy and online mass spectrometry can probe the source of hydrogen and the fate of the sacrifice reagents in photocatalytic hydrogen evolution [59]. Ultra-high spatial and temporal resolution technique, such as tip-enhanced Raman spectroscopy, can greatly improve the signal-to-noise ratio and spatial resolution, allowing for the characterization of single molecules and even single chemical bonds. Moreover, the theoretical simulation of the model systems, particularly first-principles, external field simulations, and micro-reaction dynamics simulations, is essential for exploring the fundamental mechanisms of photocatalysis. The binding energy between various intermediates and catalysts can be obtained by calculating the adsorption energy and charge density difference, further reflecting the configuration of the reaction intermediates on the catalyst surface.
Overall, ReS2 has the potential to excel as a photocatalytic material, and further investigation into its photocatalytic performance and mechanism could propel the development of photocatalysis and provide new insights into environmental and energy issues.

Author Contributions

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

Funding

This work was supported by the National Nature Science Foundation of China (22173085 and 21803061) and the Fundamental Research Funds for the Central Universities (2652019030 and 2652019031).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tian, J.; Zhao, Z.; Kumar, A.; Boughton, R.I.; Liu, H. Recent progress in design, synthesis, and applications of one-dimensional TiO2 nanostructured surface heterostructures: A review. Chem. Soc. Rev. 2014, 43, 6920–6937. [Google Scholar] [CrossRef] [PubMed]
  2. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  3. Du, J.; Lai, X.; Yang, N.; Zhai, J.; Kisailus, D.; Su, F.; Wang, D.; Jiang, L. Hierarchically Ordered Macro−Mesoporous TiO2−Graphene Composite Films: Improved Mass Transfer, Reduced Charge Recombination, and Their Enhanced Photocatalytic Activities. ACS Nano 2011, 5, 590–596. [Google Scholar] [CrossRef]
  4. Khan, M.M.; Ansari, S.A.; Pradhan, D.; Ansari, M.O.; Han, D.H.; Lee, J.; Cho, M.H. Band gap engineered TiO2 nanoparticles for visible light induced photoelectrochemical and photocatalytic studies. J. Mater. Chem. A 2014, 2, 637–644. [Google Scholar] [CrossRef]
  5. Katal, R.; Salehi, M.; Davood Abadi Farahani, M.H.; Masudy-Panah, S.; Ong, S.L.; Hu, J. Preparation of a New Type of Black TiO2 under a Vacuum Atmosphere for Sunlight Photocatalysis. ACS Appl. Mater. Interfaces 2018, 10, 35316–35326. [Google Scholar] [CrossRef] [PubMed]
  6. Yang, Y.; Ye, K.; Cao, D.; Gao, P.; Qiu, M.; Liu, L.; Yang, P. Efficient Charge Separation from F– Selective Etching and Doping of Anatase-TiO2{001} for Enhanced Photocatalytic Hydrogen Production. ACS Appl. Mater. Interfaces 2018, 10, 19633–19638. [Google Scholar] [CrossRef]
  7. Tong, H.; Ouyang, S.; Bi, Y.; Umezawa, N.; Oshikiri, M.; Ye, J. Nano-photocatalytic Materials: Possibilities and Challenges. Adv. Mater. 2012, 24, 229–251. [Google Scholar] [CrossRef]
  8. Zhang, C.; Zhou, Y.; Bao, J.; Sheng, X.; Fang, J.; Zhao, S.; Zhang, Y.; Chen, W. Hierarchical Honeycomb Br-, N-Codoped TiO2 with Enhanced Visible-Light Photocatalytic H2 Production. ACS Appl. Mater. Interfaces 2018, 10, 18796–18804. [Google Scholar] [CrossRef]
  9. Bian, Z.; Tachikawa, T.; Zhang, P.; Fujitsuka, M.; Majima, T. Au/TiO2 Superstructure-Based Plasmonic Photocatalysts Exhibiting Efficient Charge Separation and Unprecedented Activity. J. Am. Chem. Soc. 2014, 136, 458–465. [Google Scholar] [CrossRef]
  10. Yang, R.; Fan, Y.; Zhang, Y.; Mei, L.; Zhu, R.; Qin, J.; Hu, J.; Chen, Z.; Hau Ng, Y.; Voiry, D.; et al. 2D Transition Metal Dichalcogenides for Photocatalysis. Angew. Chem. Int. Ed. 2023, 62, e202218016. [Google Scholar] [CrossRef]
  11. Faraji, M.; Yousefi, M.; Yousefzadeh, S.; Zirak, M.; Naseri, N.; Jeon, T.H.; Choi, W.; Moshfegh, A.Z. Two-dimensional materials in semiconductor photoelectrocatalytic systems for water splitting. Energy Environ. Sci. 2019, 12, 59–95. [Google Scholar] [CrossRef]
  12. Tong, R.; Ng, K.W.; Wang, X.; Wang, S.; Wang, X.; Pan, H. Two-dimensional materials as novel co-catalysts for efficient solar-driven hydrogen production. J. Mater. Chem. A 2020, 8, 23202–23230. [Google Scholar] [CrossRef]
  13. Lotsch, B.V. Vertical 2D Heterostructures. Annu. Rev. Mater. Res. 2015, 45, 85–109. [Google Scholar] [CrossRef]
  14. Chhowalla, M.; Shin, H.S.; Eda, G.; Li, L.J.; Loh, K.P.; Zhang, H. The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets. Nat. Chem. 2013, 5, 263–275. [Google Scholar] [CrossRef] [PubMed]
  15. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically Thin MoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, 105, 136805. [Google Scholar] [CrossRef] [Green Version]
  16. Fu, Q.; Han, J.; Wang, X.; Xu, P.; Yao, T.; Zhong, J.; Zhong, W.; Liu, S.; Gao, T.; Zhang, Z.; et al. 2D Transition Metal Dichalcogenides: Design, Modulation, and Challenges in Electrocatalysis. Adv. Mater. 2021, 33, e1907818. [Google Scholar] [CrossRef]
  17. Pi, L.; Li, L.; Liu, K.; Zhang, Q.; Li, H.; Zhai, T. Recent Progress on 2D Noble-Transition-Metal Dichalcogenides. Adv. Funct. Mater. 2019, 29, 1904932. [Google Scholar] [CrossRef]
  18. Wu, E.; Xie, Y.; Wang, S.; Zhang, D.; Hu, X.; Liu, J. Multi-level flash memory device based on stacked anisotropic ReS2-boron nitride-graphene heterostructures. Nanoscale 2020, 12, 18800–18806. [Google Scholar] [CrossRef]
  19. Mukherjee, B.; Hayakawa, R.; Watanabe, K.; Taniguchi, T.; Nakaharai, S.; Wakayama, Y. ReS2/h-BN/Graphene Heterostructure Based Multifunctional Devices: Tunneling Diodes, FETs, Logic Gates, and Memory. Adv. Electron. Mater. 2021, 7, 2000925. [Google Scholar] [CrossRef]
  20. Qin, J.-K.; Qiu, G.; He, W.; Jian, J.; Si, M.-W.; Duan, Y.-Q.; Charnas, A.; Zemlyanov, D.Y.; Wang, H.-Y.; Shao, W.-Z.; et al. Epitaxial Growth of 1D Atomic Chain Based Se Nanoplates on Monolayer ReS2 for High-Performance Photodetectors. Adv. Funct. Mater. 2018, 28, 1806254. [Google Scholar] [CrossRef]
  21. Varghese, A.; Saha, D.; Thakar, K.; Jindal, V.; Ghosh, S.; Medhekar, N.V.; Ghosh, S.; Lodha, S. Near-Direct Bandgap WSe2/ReS2 Type-II pn Heterojunction for Enhanced Ultrafast Photodetection and High-Performance Photovoltaics. Nano Lett. 2020, 20, 1707–1717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ma, H.; Xing, Y.; Han, J.; Cui, B.; Lei, T.; Tu, H.; Guan, B.; Zeng, Z.; Zhang, B.; Lv, W. Ultrasensitive and Broad-Spectrum Photodetectors Based on InSe/ReS2 Heterostructure. Adv. Opt. Mater. 2022, 10, 2101772. [Google Scholar] [CrossRef]
  23. Su, W.; Zhang, S.; Liu, C.; Tian, Q.; Liu, X.; Li, K.; Lv, Y.; Liao, L.; Zou, X. Interlayer Transition Induced Infrared Response in ReS2/2D Perovskite van der Waals Heterostructure Photodetector. Nano Lett. 2022, 10192–10199. [Google Scholar] [CrossRef] [PubMed]
  24. Tran, T.U.; Bahng, J.; Dang, X.D.; Oh, S.; Duong, H.P.; Kang, S.S.; Yu, H.M.; Sakong, W.; Kim, M.; Choi, H.-S.; et al. Adaptive photocurrent generation of ReS2-2D Te heterostructure. Nano Energy 2022, 102, 107720. [Google Scholar] [CrossRef]
  25. Liu, Y.; Zhang, H.; Song, W.; Zhang, Y.; Hou, Z.; Zhou, G.; Zhang, Z.; Liu, J. In-situ growth of ReS2/NiS heterostructure on Ni foam as an ultra-stable electrocatalyst for alkaline hydrogen generation. Chem. Eng. J. 2023, 451, 138905. [Google Scholar] [CrossRef]
  26. Zhou, G.; Guo, Z.; Shan, Y.; Wu, S.; Zhang, J.; Yan, K.; Liu, L.; Chu, P.K.; Wu, X. High-efficiency hydrogen evolution from seawater using hetero-structured T/Td phase ReS2 nanosheets with cationic vacancies. Nano Energy 2019, 55, 42–48. [Google Scholar] [CrossRef]
  27. Ng, S.; Iffelsberger, C.; Sofer, Z.; Pumera, M. Tunable Room-Temperature Synthesis of ReS2 Bicatalyst on 3D- and 2D-Printed Electrodes for Photo- and Electrochemical Energy Applications. Adv. Funct. Mater. 2020, 30, 1910193. [Google Scholar] [CrossRef]
  28. Cheng, P.; Liu, Y.; Ziegler, M.; Klingenhof, M.; Wang, D.; Zhang, Z.; Strasser, P.; Schaaf, P. Improving Silicon Photocathode Performance for Water Reduction through Dual Interface Engineering and Integrating ReS2 Photocatalyst. ACS Appl. Energy Mater. 2022, 5, 8222–8231. [Google Scholar] [CrossRef]
  29. Tongay, S.; Sahin, H.; Ko, C.; Luce, A.; Fan, W.; Liu, K.; Zhou, J.; Huang, Y.-S.; Ho, C.-H.; Yan, J.; et al. Monolayer behaviour in bulk ReS2 due to electronic and vibrational decoupling. Nat. Commun. 2014, 5, 3252. [Google Scholar] [CrossRef] [Green Version]
  30. Lin, Y.-C.; Komsa, H.-P.; Yeh, C.-H.; Björkman, T.; Liang, Z.-Y.; Ho, C.-H.; Huang, Y.-S.; Chiu, P.-W.; Krasheninnikov, A.V.; Suenaga, K. Single-Layer ReS2: Two-Dimensional Semiconductor with Tunable In-Plane Anisotropy. ACS Nano 2015, 9, 11249–11257. [Google Scholar] [CrossRef]
  31. Feng, Y.; Zhou, W.; Wang, Y.; Zhou, J.; Liu, E.; Fu, Y.; Ni, Z.; Wu, X.; Yuan, H.; Miao, F.; et al. Raman vibrational spectra of bulk to monolayer ReS2 with lower symmetry. Phys. Rev. B 2015, 92, 054110. [Google Scholar] [CrossRef] [Green Version]
  32. Rahman, M.; Davey, K.; Qiao, S.Z. Advent of 2D Rhenium Disulfide (ReS2): Fundamentals to Applications. Adv. Funct. Mater. 2017, 27, 1606129. [Google Scholar] [CrossRef] [Green Version]
  33. Zhang, Q.; Wang, W.; Zhang, J.; Zhu, X.; Zhang, Q.; Zhang, Y.; Ren, Z.; Song, S.; Wang, J.; Ying, Z.; et al. Highly Efficient Photocatalytic Hydrogen Evolution by ReS2 via a Two-Electron Catalytic Reaction. Adv. Mater. 2018, 30, e1707123. [Google Scholar] [CrossRef] [PubMed]
  34. Ran, J.; Zhang, H.; Qu, J.; Shan, J.; Chen, S.; Yang, F.; Zheng, R.; Cairney, J.; Song, L.; Jing, L.; et al. Atomic-Level Insights into the Edge Active ReS2 Ultrathin Nanosheets for High-Efficiency Light-to-Hydrogen Conversion. ACS Mater. Lett. 2020, 2, 1484–1494. [Google Scholar] [CrossRef]
  35. Zhang, Q.; Fu, L. Novel Insights and Perspectives into Weakly Coupled ReS2 toward Emerging Applications. Chem 2019, 5, 505–525. [Google Scholar] [CrossRef] [Green Version]
  36. Cao, Y.-D.; Sun, Y.-H.; Shi, S.-F.; Wang, R.-M. Anisotropy of two-dimensional ReS2 and advances in its device application. Rare Met. 2021, 40, 3357–3374. [Google Scholar] [CrossRef]
  37. Xie, X.; Mao, M.; Qi, S.; Ma, J. ReS2-Based electrode materials for alkali-metal ion batteries. CrystEngComm 2019, 21, 3755–3769. [Google Scholar] [CrossRef]
  38. Xiong, Y.; Chen, H.; Zhang, D.W.; Zhou, P. Electronic and Optoelectronic Applications Based on ReS2. Phys. Status Solidi Rapid Res. Lett. 2019, 13, 1800658. [Google Scholar] [CrossRef]
  39. Fadhel, M.M.; Ali, N.; Rashid, H.; Sapiee, N.M.; Hamzah, A.E.; Zan, M.S.D.; Aziz, N.A.; Arsad, N. A Review on Rhenium Disulfide: Synthesis Approaches, Optical Properties, and Applications in Pulsed Lasers. Nanomaterials 2021, 11, 2367. [Google Scholar] [CrossRef]
  40. Ren, H.; Xiang, G. Recent Progress in Research on Ferromagnetic Rhenium Disulfide. Nanomaterials 2022, 12, 3451. [Google Scholar] [CrossRef]
  41. Wang, X.; Wang, X.; Huang, J.; Li, S.; Meng, A.; Li, Z. Interfacial chemical bond and internal electric field modulated Z-scheme Sv-ZnIn2S4/MoSe2 photocatalyst for efficient hydrogen evolution. Nat. Commun. 2021, 12, 4112. [Google Scholar] [CrossRef] [PubMed]
  42. Zhao, D.; Wang, Y.; Dong, C.-L.; Huang, Y.-C.; Chen, J.; Xue, F.; Shen, S.; Guo, L. Boron-doped nitrogen-deficient carbon nitride-based Z-scheme heterostructures for photocatalytic overall water splitting. Nat. Energy 2021, 6, 388–397. [Google Scholar] [CrossRef]
  43. Xi, Y.; Wang, F.; Feng, H.; Xiong, Y.; Huang, Q. Cauliflower-like Mn0.2Cd0.8S decorated with ReS2 nanosheets for boosting photocatalytic H2 evolution activity. New J. Chem. 2021, 45, 15949–15955. [Google Scholar] [CrossRef]
  44. Ma, X.; Cheng, H. ReS2 with unique trion behavior as a co-catalyst for enhanced sunlight hydrogen production. J. Colloid Interface Sci. 2022, 634, 32–43. [Google Scholar] [CrossRef] [PubMed]
  45. Liu, J.; Chen, K.; Pan, G.M.; Luo, Z.J.; Xie, Y.; Li, Y.Y.; Lin, Y.J.; Hao, Z.H.; Zhou, L.; Ding, S.J.; et al. Largely enhanced photocatalytic hydrogen production rate of CdS/(Au-ReS2) nanospheres by the dielectric-plasmon hybrid antenna effect. Nanoscale 2018, 10, 19586–19594. [Google Scholar] [CrossRef]
  46. Xing, C.; Zhao, H.; Yu, G.; Guo, L.; Hu, Y.; Chen, T.; Jiang, L.; Li, X. Modification of g-C3N4 Photocatalyst with Flower-like ReS2 for Highly Efficient Photocatalytic Hydrogen Evolution. ChemCatChem 2020, 12, 6385–6392. [Google Scholar] [CrossRef]
  47. Jing, Q.; Zhang, H.; Huang, H.; Fan, X.; Zhang, Y.; Hou, X.; Xu, Q.; Ni, Z.; Qiu, T. Ultrasonic exfoliated ReS2 nanosheets: Fabrication and use as co-catalyst for enhancing photocatalytic efficiency of TiO2 nanoparticles under sunlight. Nanotechnology 2019, 30, 184001. [Google Scholar] [CrossRef] [PubMed]
  48. Fujita, T.; Ito, Y.; Tan, Y.; Yamaguchi, H.; Hojo, D.; Hirata, A.; Voiry, D.; Chhowalla, M.; Chen, M. Chemically exfoliated ReS2 nanosheets. Nanoscale 2014, 6, 12458–12462. [Google Scholar] [CrossRef]
  49. Yu, Z.G.; Cai, Y.; Zhang, Y.-W. Robust Direct Bandgap Characteristics of One- and Two-Dimensional ReS2. Sci. Rep. 2015, 5, 13783. [Google Scholar] [CrossRef] [Green Version]
  50. Ran, J.; Zhang, H.; Qu, J.; Shan, J.; Davey, K.; Cairney, J.M.; Jing, L.; Qiao, S.Z. Significantly Raised Visible-Light Photocatalytic H2 Evolution on a 2D/2D ReS2 /In2ZnS4 van der Waals Heterostructure. Small 2021, 17, 2100296. [Google Scholar] [CrossRef]
  51. Zhang, Y.; Yao, D.; Xia, B.; Xu, H.; Tang, Y.; Davey, K.; Ran, J.; Qiao, S.-Z. ReS2 Nanosheets with In Situ Formed Sulfur Vacancies for Efficient and Highly Selective Photocatalytic CO2 Reduction. Small Sci. 2021, 1, 2000052. [Google Scholar] [CrossRef]
  52. Liu, N.; Dai, W.; Fei, F.; Xu, H.; Lei, J.; Quan, G.; Zheng, Y.; Zhang, X.; Tang, L. Insights into the photocatalytic activation persulfate by visible light over ReS2/MIL-88B(Fe) for highly efficient degradation of ibuprofen: Combination of experimental and theoretical study. Sep. Purif. Technol. 2022, 297, 121545. [Google Scholar] [CrossRef]
  53. Zhan, X.; Ou, D.; Zheng, Y.; Li, B.; Xu, L.; Yang, H.; Yang, W.; Zhang, H.; Hou, H.; Yang, W. Boosted photocatalytic hydrogen production over two-dimensional/two-dimensional Ta3N5/ReS2 van der Waals heterojunctions. J. Colloid Interface Sci. 2023, 629, 455–466. [Google Scholar] [CrossRef] [PubMed]
  54. Yang, T.; Shao, Y.; Hu, J.; Qu, J.; Yang, X.; Yang, F.; Ming Li, C. Ultrathin layered 2D/2D heterojunction of ReS2/high-crystalline g-C3N4 for significantly improved photocatalytic hydrogen evolution. Chem. Eng. J. 2022, 448, 137613. [Google Scholar] [CrossRef]
  55. Knirsch, K.C.; Berner, N.C.; Nerl, H.C.; Cucinotta, C.S.; Gholamvand, Z.; McEvoy, N.; Wang, Z.; Abramovic, I.; Vecera, P.; Halik, M.; et al. Basal-Plane Functionalization of Chemically Exfoliated Molybdenum Disulfide by Diazonium Salts. ACS Nano 2015, 9, 6018–6030. [Google Scholar] [CrossRef] [PubMed]
  56. Yu, J.; Seo, S.; Luo, Y.; Sun, Y.; Oh, S.; Nguyen, C.T.K.; Seo, C.; Kim, J.H.; Kim, J.; Lee, H. Efficient and Stable Solar Hydrogen Generation of Hydrophilic Rhenium-Disulfide-Based Photocatalysts via Chemically Controlled Charge Transfer Paths. ACS Nano 2020, 14, 1715–1726. [Google Scholar] [CrossRef]
  57. Wang, X.; Xue, Y.; Liang, Z.; Tian, J.; Zhang, X.; Chen, X. Insights into the function of semi-metallic 1T’ phase ReS2 as cocatalyst decorated g-C3N4 nanotubes for enhanced photocatalytic hydrogen production activity. Mater. Today Adv. 2022, 15, 100257. [Google Scholar] [CrossRef]
  58. Wang, X.; Chen, B.; Yan, D.; Zhao, X.; Wang, C.; Liu, E.; Zhao, N.; He, F. Distorted 1T-ReS2 Nanosheets Anchored on Porous TiO2 Nanofibers for Highly Enhanced Photocatalytic Hydrogen Production. ACS Appl. Mater. Interfaces 2019, 11, 23144–23151. [Google Scholar] [CrossRef]
  59. Ye, L.; Ma, Z.; Deng, Y.; Ye, Y.; Li, W.; Kou, M.; Xie, H.; Zhikun, X.; Zhou, Y.; Xia, D.; et al. Robust and efficient photocatalytic hydrogen generation of ReS2/CdS and mechanistic study by on-line mass spectrometry and in situ infrared spectroscopy. Appl. Catal. B 2019, 257, 117897. [Google Scholar] [CrossRef]
  60. Zhan, X.; Fang, Z.; Li, B.; Zhang, H.; Xu, L.; Hou, H.; Yang, W. Rationally designed Ta3N5@ReS2 heterojunctions for promoted photocatalytic hydrogen production. J. Mater. Chem. A 2021, 9, 27084–27094. [Google Scholar] [CrossRef]
  61. Xiong, X.; Yan, A.; Zhang, X.; Huang, F.; Li, Z.; Zhang, Z.; Weng, H. ReS2/ZnIn2S4 heterojunctions with enhanced visible-light-driven hydrogen evolution performance for water splitting. J. Alloys Compd. 2021, 873, 159850. [Google Scholar] [CrossRef]
  62. Chen, R.; Ma, M.; Luo, Y.; Qian, L.; Xu, S. Enhanced interfacial effect between CdS and ReS2 on boosted hydrogen evolution performance via phase structure engineering. J. Solid State Chem. 2022, 312, 123238. [Google Scholar] [CrossRef]
  63. Song, T.; Wang, J.; Su, L.; Xu, H.; Bai, X.; Zhou, L.; Tu, W. Promotion effect of rhenium on MoS2/ReS2@CdS nanostructures for photocatalytic hydrogen production. Mol. Catal. 2021, 516, 111939. [Google Scholar] [CrossRef]
  64. Liu, J.; Qi, F.; Zhang, N.; Yang, J.; Liang, Z.; Tian, C.; Zhang, W.; Tang, X.; Wu, D.; Huang, Q. Three-dimensional structural ReS2@Cu2O/Cu heterojunction photocatalysts for visible-light-driven CO2 reduction. J. Mater. Sci. 2022, 57, 15474–15487. [Google Scholar] [CrossRef]
  65. Zhang, L.; Zheng, Q.; Xie, Y.; Lan, Z.; Prezhdo, O.V.; Saidi, W.A.; Zhao, J. Delocalized Impurity Phonon Induced Electron-Hole Recombination in Doped Semiconductors. Nano Lett. 2018, 18, 1592–1599. [Google Scholar] [CrossRef] [PubMed]
  66. Ji, Q.; Zhang, Y.; Zhang, Y.; Liu, Z. Chemical vapour deposition of group-VIB metal dichalcogenide monolayers: Engineered substrates from amorphous to single crystalline. Chem. Soc. Rev. 2015, 44, 2587–2602. [Google Scholar] [CrossRef] [PubMed]
  67. Shi, Y.; Li, H.; Li, L.-J. Recent advances in controlled synthesis of two-dimensional transition metal dichalcogenides via vapour deposition techniques. Chem. Soc. Rev. 2015, 44, 2744–2756. [Google Scholar] [CrossRef]
  68. Zereshki, P.; Yao, P.; He, D.; Wang, Y.; Zhao, H. Interlayer charge transfer in ReS2/WS2 van der Waals heterostructures. Phys. Rev. B 2019, 99, 195438. [Google Scholar] [CrossRef]
  69. Varghese, A.; Joseph, D.; Ghosh, S.; Thakar, K.; Mcdhckar, N.; Lodha, S. WSe2/ReS2 vdW Heterostructure for Versatile Optoelectronic Applications. In Proceedings of the 2018 76th Device Research Conference (DRC), Santa Barbara, CA, USA, 24–27 June 2018. [Google Scholar]
  70. Wei, N.; Cai, J.; Wang, R.; Wang, M.; Lv, W.; Ci, H.; Sun, J.; Liu, Z. Elevated polysulfide regulation by an ultralight all-CVD-built ReS2@N-Doped graphene heterostructure interlayer for lithium-sulfur batteries. Nano Energy. 2019, 66, 104190. [Google Scholar]
  71. Li, J.; Song, P.; Zhao, J.; Vaklinova, K.; Zhao, X.; Li, Z.; Qiu, Z.; Wang, Z.; Lin, L.; Zhao, M.; et al. Printable two-dimensional superconducting monolayers. Nat. Mater. 2021, 20, 181–187. [Google Scholar]
  72. Liu, D.; Hong, J.; Li, X.; Zhou, X.; Jin, B.; Cui, Q.; Chen, J.; Feng, Q.; Xu, C.; Zhai, T.; et al. Synthesis of 2H-1T′ WS2-ReS2 Heterophase Structures with Atomically Sharp Interface via Hydrogen-Triggered One-Pot Growth. Adv. Funct. Mater. 2020, 30, 1910169. [Google Scholar] [CrossRef]
  73. Chen, B.; Wu, K.; Suslu, A.; Yang, S.; Cai, H.; Yano, A.; Soignard, E.; Aoki, T.; March, K.; Shen, Y.; et al. Controlling Structural Anisotropy of Anisotropic 2D Layers in Pseudo-1D/2D Material Heterojunctions. Adv. Mater. 2017, 29, 1701201. [Google Scholar] [CrossRef] [PubMed]
  74. Liu, D.; Hong, J.; Wang, X.; Li, X.; Feng, Q.; Tan, C.; Zhai, T.; Ding, F.; Peng, H.; Xu, H. Diverse Atomically Sharp Interfaces and Linear Dichroism of 1T’ ReS2-ReSe2 Lateral p–n Heterojunctions. Adv. Funct. Mater. 2018, 28, 1804696. [Google Scholar] [CrossRef]
  75. Zhang, T.; Jiang, B.; Xu, Z.; Mendes, R.G.; Xiao, Y.; Chen, L.; Fang, L.; Gemming, T.; Chen, S.; Rummeli, M.H.; et al. Twinned growth behaviour of two-dimensional materials. Nat. Commun. 2016, 7, 13911. [Google Scholar] [CrossRef] [Green Version]
  76. Lim, J.; Kadyrov, A.; Jeon, D.; Choi, Y.; Bae, J.; Lee, S. Contact Engineering of Vertically Grown ReS2 with Schottky Barrier Modulation. ACS Appl. Mater. Interfaces 2021, 13, 7529–7538. [Google Scholar] [CrossRef]
  77. Zhou, J.; Lin, J.; Huang, X.; Zhou, Y.; Chen, Y.; Xia, J.; Wang, H.; Xie, Y.; Yu, H.; Lei, J.; et al. A library of atomically thin metal chalcogenides. Nature 2018, 556, 355–359. [Google Scholar] [CrossRef]
  78. Ghoshal, D.; Yoshimura, A.; Gupta, T.; House, A.; Basu, S.; Chen, Y.; Wang, T.; Yang, Y.; Shou, W.; Hachtel, J.A.; et al. Theoretical and Experimental Insight into the Mechanism for Spontaneous Vertical Growth of ReS2 Nanosheets. Adv. Funct. Mater. 2018, 28, 1801286. [Google Scholar] [CrossRef]
  79. Zhang, H.; Wang, S.; Yu, J. Low-Symmetry Two-Dimensional ReS2 and its Heterostructures: Chemical Vapor Deposition Synthesis and Properties. Prog. Chem. 2022, 34, 1440–1452. [Google Scholar]
  80. Gao, J.; Li, L.; Tan, J.; Sun, H.; Li, B.; Idrobo, J.C.; Singh, C.V.; Lu, T.M.; Koratkar, N. Vertically Oriented Arrays of ReS2 Nanosheets for Electrochemical Energy Storage and Electrocatalysis. Nano Lett. 2016, 16, 3780–3787. [Google Scholar] [CrossRef]
  81. Kang, Y.B.; Han, X.; Kim, S.; Yuan, H.; Ling, N.; Ham, H.C.; Dai, L.; Park, H.S. Structural Engineering of Ultrathin ReS2 on Hierarchically Architectured Graphene for Enhanced Oxygen Reduction. ACS Nano 2021, 15, 5560–5566. [Google Scholar] [CrossRef]
  82. He, X.; Liu, F.; Hu, P.; Fu, W.; Wang, X.; Zeng, Q.; Zhao, W.; Liu, Z. Chemical Vapor Deposition of High-Quality and Atomically Layered ReS2. Small 2015, 11, 5423–5429. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, L.; Chu, H.; Zhang, X.; Pan, H.; Zhao, S.; Li, D. Heterostructure ReS2/GaAs Saturable Absorber Passively Q-Switched Nd:YVO4 Laser. Nanoscale Res. Lett. 2019, 14, 112. [Google Scholar] [CrossRef] [PubMed]
  84. Butanovs, E.; Kuzmin, A.; Piskunov, S.; Smits, K.; Kalinko, A.; Polyakov, B. Synthesis and characterization of GaN/ReS2, ZnS/ReS2 and ZnO/ReS2 core/shell nanowire heterostructures. Appl. Surf. Sci. 2021, 536, 147841. [Google Scholar] [CrossRef]
  85. Zhang, Y.; Liu, Y.; Cheng, P.; Song, W.; Zhang, X.; Rong, S.; Gao, X.; Zhou, G.; Zhang, Z.; Liu, J. Universal substrate growth of Ag-modified ReS2 as visible-light-driven photocatalyst for highly efficient water disinfection. Chem. Eng. J. 2022, 430, 132918. [Google Scholar] [CrossRef]
  86. Zhou, G.; Wu, Q.; Wu, L.; Liu, L.; Wang, D.; Wang, P. Reaction kinetic acceleration induced by atomic-hybridized channels in carbon quantum dot/ReS2 composites for efficient Cr(VI) reduction. Appl. Catal. B 2022, 300, 119807. [Google Scholar] [CrossRef]
  87. Zhao, Y.; Song, W.; Li, Z.; Zhang, Z.; Zhou, G. A new strategy: Fermi level control to realize 3D pyramidal NiCo-LDH/ReS2/n-PSi as a high-performance photoanode for the oxygen evolution reaction. J. Mater. Chem. C 2022, 10, 3848–3855. [Google Scholar] [CrossRef]
  88. Ha, E.; Liu, W.; Wang, L.; Man, H.W.; Hu, L.; Tsang, S.C.; Chan, C.T.; Kwok, W.M.; Lee, L.Y.; Wong, K.Y. Cu2ZnSnS4/MoS2-Reduced Graphene Oxide Heterostructure: Nanoscale Interfacial Contact and Enhanced Photocatalytic Hydrogen Generation. Sci. Rep. 2017, 7, 39411. [Google Scholar] [CrossRef] [Green Version]
  89. Li, L.; Salvador, P.A.; Rohrer, G.S. Photocatalysts with internal electric fields. Nanoscale 2014, 6, 24–42. [Google Scholar] [CrossRef] [Green Version]
  90. Wu, Y.; Wang, H.; Sun, Y.; Xiao, T.; Tu, W.; Yuan, X.; Zeng, G.; Li, S.; Chew, J.W. Photogenerated charge transfer via interfacial internal electric field for significantly improved photocatalysis in direct Z-scheme oxygen-doped carbon nitrogen/CoAl-layered double hydroxide heterojunction. Appl. Catal. B 2018, 227, 530–540. [Google Scholar] [CrossRef]
  91. Bellus, M.Z.; Li, M.; Lane, S.D.; Ceballos, F.; Cui, Q.; Zeng, X.C.; Zhao, H. Type-I van der Waals heterostructure formed by MoS2 and ReS2 monolayers. Nanoscale Horiz. 2017, 2, 31–36. [Google Scholar] [CrossRef]
  92. Wang, F.; Hu, J.; Liang, R.; Lei, W.; Lou, Z.; Pan, X.; Lu, B.; Ye, Z. Novel ReS2/g-C3N4 heterojunction photocatalyst formed by electrostatic self-assembly with increased H2 production. Int. J. Hydrogen Energy 2022, 47, 29284–29294. [Google Scholar] [CrossRef]
  93. Hu, J.; Li, X.; Qu, J.; Yang, X.; Cai, Y.; Yang, T.; Yang, F.; Li, C.M. Bifunctional honeycomb hierarchical structured 3D/3D ReS2/ZnIn2S4-Sv heterojunction for efficient photocatalytic H2-evolution integrated with biomass oxidation. Chem. Eng. J. 2023, 453, 139957. [Google Scholar] [CrossRef]
  94. Reza Gholipour, M.; Dinh, C.T.; Beland, F.; Do, T.O. Nanocomposite heterojunctions as sunlight-driven photocatalysts for hydrogen production from water splitting. Nanoscale 2015, 7, 8187–8208. [Google Scholar] [CrossRef] [PubMed]
  95. Saeed, M.; Uddin, W.; Saleemi, A.S.; Hafeez, M.; Kamil, M.; Mir, I.A.; Sunila; Ullah, R.; Rehman, S.U.; Ling, Z. Optoelectronic properties of MoS2-ReS2 and ReS2-MoS2 heterostructures. Phys. B 2020, 577, 411809. [Google Scholar] [CrossRef]
  96. Zhang, Y.; Yan, Y.; Shen, G.; Hong, K. Synthesis of type-II heterojunction films between ReS2 and XS2 (X = Mo, W) with high electrocatalystic activities in dye-sensitized solar cells. Catal. Commun. 2022, 170, 106497. [Google Scholar] [CrossRef]
  97. Zhang, W.; Mohamed, A.R.; Ong, W.-J. Z-Scheme Photocatalytic Systems for Carbon Dioxide Reduction: Where Are We Now? Angew. Chem. Int. Ed. 2020, 59, 22894–22915. [Google Scholar] [CrossRef]
  98. Abdul Nasir, J.; Munir, A.; Ahmad, N.; ul Haq, T.; Khan, Z.; Rehman, Z. Photocatalytic Z-Scheme Overall Water Splitting: Recent Advances in Theory and Experiments. Adv. Mater. 2021, 33, 2105195. [Google Scholar] [CrossRef]
  99. Wang, L.; Bie, C.; Yu, J. Challenges of Z-scheme photocatalytic mechanisms. Trends Chem. 2022, 4, 973–983. [Google Scholar] [CrossRef]
  100. Di, T.; Xu, Q.; Ho, W.; Tang, H.; Xiang, Q.; Yu, J. Review on Metal Sulphide-based Z-scheme Photocatalysts. ChemCatChem 2019, 11, 1394–1411. [Google Scholar] [CrossRef]
  101. Zhang, G.; Wang, Z.; Wu, J. Construction of a Z-scheme heterojunction for high-efficiency visible-light-driven photocatalytic CO2 reduction. Nanoscale 2021, 13, 4359–4389. [Google Scholar] [CrossRef]
  102. Liu, H.; Xu, B.; Liu, J.M.; Yin, J.; Miao, F.; Duan, C.G.; Wan, X.G. Highly efficient and ultrastable visible-light photocatalytic water splitting over ReS2. Phys. Chem. Chem. Phys. 2016, 18, 14222–14227. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, Y.; Shi, R.; Song, K.; Liu, C.; He, F. Constructing a 2D/2D interfacial contact in ReS2/TiO2 via Ti–S bond for efficient charge transfer in photocatalytic hydrogen production. J. Mater. Chem. A 2021, 9, 23687–23696. [Google Scholar] [CrossRef]
  104. Lin, B.; Ma, B.; Chen, J.; Zhou, Y.; Zhou, J.; Yan, X.; Xue, C.; Luo, X.; Liu, Q.; Wang, J.; et al. Sea-urchin-like ReS2 nanosheets with charge edge-collection effect as a novel cocatalyst for high-efficiency photocatalytic H2 evolution. Chin. Chem. Lett. 2022, 33, 943–947. [Google Scholar] [CrossRef]
  105. Guo, L.; Yu, G.; Zhao, H.; Xing, C.; Hu, Y.; Chen, T.; Li, X. Construction of heterojunctions between ReS2 and twin crystal ZnxCd1−xS for boosting solar hydrogen evolution. New J. Chem. 2021, 45, 5137–5145. [Google Scholar] [CrossRef]
  106. Shi, S.; Sun, Z.; Hu, Y.H. Synthesis, stabilization and applications of 2-dimensional 1T metallic MoS2. J. Mater. Chem. A 2018, 6, 23932–23977. [Google Scholar] [CrossRef]
  107. Kwak, I.H.; Kwon, I.S.; Debela, T.T.; Seo, J.; Ahn, J.-P.; Yoo, S.J.; Kim, J.-G.; Park, J.; Kang, H.S. Two-dimensional MoS2–melamine hybrid nanostructures for enhanced catalytic hydrogen evolution reaction. J. Mater. Chem. A 2019, 7, 22571–22578. [Google Scholar] [CrossRef]
  108. Liu, L.; Wu, J.; Wu, L.; Ye, M.; Liu, X.; Wang, Q.; Hou, S.; Lu, P.; Sun, L.; Zheng, J.; et al. Phase-selective synthesis of 1T’ MoS2 monolayers and heterophase bilayers. Nat. Mater. 2018, 17, 1108–1114. [Google Scholar] [CrossRef]
  109. Yu, Y.; Nam, G.H.; He, Q.; Wu, X.J.; Zhang, K.; Yang, Z.; Chen, J.; Ma, Q.; Zhao, M.; Liu, Z.; et al. High phase-purity 1T’-MoS2- and 1T’-MoSe2-layered crystals. Nat. Chem. 2018, 10, 638–643. [Google Scholar] [CrossRef]
  110. Ran, J.; Chen, L.; Wang, D.; Talebian-Kiakalaieh, A.; Jiao, Y.; Adel Hamza, M.; Qu, Y.; Jing, L.; Davey, K.; Qiao, S.-Z. Atomic-Level Regulated Two-Dimensional ReSe2: A Universal Platform Boosting Photocatalysis. Adv. Mater. 2023, e2210164. [Google Scholar] [CrossRef]
  111. Huang, Q.; Yang, J.; Qi, F.; Zhang, W.; Zhang, N.; Liang, Z.; Liu, J.; Tian, C.; Tang, X.; Wu, D.; et al. Visible light driven photocatalytic reduction of CO2 on Au-Pt/Cu2O/ReS2 with high efficiency and controllable selectivity. Chem. Eng. J. 2022, 437, 135299. [Google Scholar] [CrossRef]
  112. Li, X.; Wang, X.; Hong, J.; Liu, D.; Feng, Q.; Lei, Z.; Liu, K.; Ding, F.; Xu, H. Nanoassembly Growth Model for Subdomain and Grain Boundary Formation in 1T′ Layered ReS2. Adv. Funct. Mater. 2019, 29, 1906385. [Google Scholar] [CrossRef]
  113. Ho, C.H.; Huang, Y.S.; Tiong, K.K. In-plane anisotropy of the optical and electrical properties of ReS2 and ReSe2 layered crystals. J. Alloys Compd. 2001, 317–318, 222–226. [Google Scholar] [CrossRef]
  114. Parasuraman, P.S.; Ho, J.-H.; Lin, M.-H.; Ho, C.-H. In-Plane Axially Enhanced Photocatalysis by Re4 Diamond Chains in Layered ReS2. J. Phys. Chem. C 2018, 122, 18776–18784. [Google Scholar] [CrossRef]
  115. Liu, W.; Wang, P.; Ao, Y.; Chen, J.; Gao, X.; Jia, B.; Ma, T. Directing Charge Transfer in a Chemical-Bonded BaTiO3 @ReS2 Schottky Heterojunction for Piezoelectric Enhanced Photocatalysis. Adv. Mater. 2022, 34, e2202508. [Google Scholar] [CrossRef]
  116. Li, H.; Liang, Z.; Deng, Q.; Hu, M.T.; Du, N.; Hou, W. Facile Construction of Defect-rich Rhenium Disulfide/Graphite Carbon Nitride Heterojunction via Electrostatic Assembly for Fast Charge Separation and Photoactivity Enhancement. ChemCatChem 2019, 11, 1633–1642. [Google Scholar] [CrossRef]
  117. Tu, S.; Guo, Y.; Zhang, Y.; Hu, C.; Zhang, T.; Ma, T.; Huang, H. Piezocatalysis and Piezo-Photocatalysis: Catalysts Classification and Modification Strategy, Reaction Mechanism, and Practical Application. Adv. Funct. Mater. 2020, 30, 2005158. [Google Scholar] [CrossRef]
  118. Nosaka, Y.; Nosaka, A.Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef]
  119. Wang, W.; Huang, G.; Yu, J.C.; Wong, P.K. Advances in photocatalytic disinfection of bacteria: Development of photocatalysts and mechanisms. J. Environ. Sci. 2015, 34, 232–247. [Google Scholar] [CrossRef]
Figure 1. Top view and side view of (a) MoS2 and (b) ReS2. (c) The electronic band structure of bulk (orange solid curves) and monolayer (purple dashed curves) ReS2, as calculated by DFT, is predicted to be a direct bandgap semiconductor with a nearly identical bandgap value at the Γ point [29]. Reproduced with permission: Copyright 2014, Nature Publishing Group.
Figure 1. Top view and side view of (a) MoS2 and (b) ReS2. (c) The electronic band structure of bulk (orange solid curves) and monolayer (purple dashed curves) ReS2, as calculated by DFT, is predicted to be a direct bandgap semiconductor with a nearly identical bandgap value at the Γ point [29]. Reproduced with permission: Copyright 2014, Nature Publishing Group.
Molecules 28 02627 g001
Figure 2. (a) Liquid exfoliation of ReS2 nanosheets from bulk material [47]. Reproduced with permission: Copyright 2019, IOP Publishing. Atomic-resolution HAADF-STEM images of (b) basal plane and (c) edge of ReS2 nanosheets [50]. (d) AFM image and (e) thickness profiles of ReS2 nanosheets [54]. Reproduced with permission: Copyright 2021, Wiley-VCH Verlag GmbH & Co. KGaA. (f) Schematic illustration of the solvothermal synthesis process of g-C3N4 nanotubes and 1T’-ReS2/g-C3N4 composites [57]. Reproduced with permission: Copyright 2022, ELSEVIER. (g) XRD patterns of 1T’-ReS2 and 2H-ReS2 after heat treatment [57]. (h) TEM images of 1T’-ReS2/g-C3N4 composite (12 wt%) [57]. Reproduced with permission: Copyright 2022, ELSEVIER. (i) Schematic diagram of the CVD-synthesized ReS2 and the two-electron photocatalytic reaction [33]. (j) SEM image of the lateral view of the ReS2 nanowalls on the silicon substrate; scale bar: 200 nm [33]. (k) HRTEM characterization of a ReS2 NW with clear lattice fringe; scale bar: 2 nm [33]. Reproduced with permission: Copyright 2018, Wiley-VCH Verlag GmbH & Co. KGaA.
Figure 2. (a) Liquid exfoliation of ReS2 nanosheets from bulk material [47]. Reproduced with permission: Copyright 2019, IOP Publishing. Atomic-resolution HAADF-STEM images of (b) basal plane and (c) edge of ReS2 nanosheets [50]. (d) AFM image and (e) thickness profiles of ReS2 nanosheets [54]. Reproduced with permission: Copyright 2021, Wiley-VCH Verlag GmbH & Co. KGaA. (f) Schematic illustration of the solvothermal synthesis process of g-C3N4 nanotubes and 1T’-ReS2/g-C3N4 composites [57]. Reproduced with permission: Copyright 2022, ELSEVIER. (g) XRD patterns of 1T’-ReS2 and 2H-ReS2 after heat treatment [57]. (h) TEM images of 1T’-ReS2/g-C3N4 composite (12 wt%) [57]. Reproduced with permission: Copyright 2022, ELSEVIER. (i) Schematic diagram of the CVD-synthesized ReS2 and the two-electron photocatalytic reaction [33]. (j) SEM image of the lateral view of the ReS2 nanowalls on the silicon substrate; scale bar: 200 nm [33]. (k) HRTEM characterization of a ReS2 NW with clear lattice fringe; scale bar: 2 nm [33]. Reproduced with permission: Copyright 2018, Wiley-VCH Verlag GmbH & Co. KGaA.
Molecules 28 02627 g002
Figure 3. Diagram of (a) Type-I, (b) Type-II, and (c) all-solid-state Z-Scheme heterojunction. (d) Direct Z-scheme heterojunction.
Figure 3. Diagram of (a) Type-I, (b) Type-II, and (c) all-solid-state Z-Scheme heterojunction. (d) Direct Z-scheme heterojunction.
Molecules 28 02627 g003
Figure 4. (a) The formation procedure of ReS2 nanosheet/TiO2 nanoparticle heterojunctions (SURTHs). (b) TEM images of SURTHs (c) Schematic of the sea-urchin-like cocatalyst with the charge edge-collection effect loaded on the photocatalyst for H2 evolution. (d) Time-dependent photocatalytic hydrogen production of Bulk ReS2/TiO2 and SURTHs under simulated solar-light irradiation. (e) Cycling tests for photocatalytic hydrogen production over SURTHs under simulated solar-light irradiation [104]. Reproduced with permission: Copyright 2022, ELSEVIER.
Figure 4. (a) The formation procedure of ReS2 nanosheet/TiO2 nanoparticle heterojunctions (SURTHs). (b) TEM images of SURTHs (c) Schematic of the sea-urchin-like cocatalyst with the charge edge-collection effect loaded on the photocatalyst for H2 evolution. (d) Time-dependent photocatalytic hydrogen production of Bulk ReS2/TiO2 and SURTHs under simulated solar-light irradiation. (e) Cycling tests for photocatalytic hydrogen production over SURTHs under simulated solar-light irradiation [104]. Reproduced with permission: Copyright 2022, ELSEVIER.
Molecules 28 02627 g004
Figure 5. (a) Preparation for multiple junctions between functionalized ReS2 nanosheets and two-phase TiO2. (b) Proposed transfer paths of photogenerated charges in terms of contact properties at the interface. Type II: The energy bands of TiO2 and nonfunctionalized ReS2 nanosheet can bend upward and downward toward the physically contacted interface, respectively. Z-scheme: The energy bands of TiO2 and ReS2−C6H5COOH can bend downward and upward toward the chemically contacted interface, respectively. (c,d) Proposed energy band diagrams for ReS2 NS/TiO2 (type II) and ReS2−BzO−TiO2 (Z-scheme) based on the flat-band potentials of each component (EC—conduction band energy; EV—valence band energy; Efb—flat-band energy) [56]. Reproduced with permission: Copyright 2020, American Chemical Society.
Figure 5. (a) Preparation for multiple junctions between functionalized ReS2 nanosheets and two-phase TiO2. (b) Proposed transfer paths of photogenerated charges in terms of contact properties at the interface. Type II: The energy bands of TiO2 and nonfunctionalized ReS2 nanosheet can bend upward and downward toward the physically contacted interface, respectively. Z-scheme: The energy bands of TiO2 and ReS2−C6H5COOH can bend downward and upward toward the chemically contacted interface, respectively. (c,d) Proposed energy band diagrams for ReS2 NS/TiO2 (type II) and ReS2−BzO−TiO2 (Z-scheme) based on the flat-band potentials of each component (EC—conduction band energy; EV—valence band energy; Efb—flat-band energy) [56]. Reproduced with permission: Copyright 2020, American Chemical Society.
Molecules 28 02627 g005
Figure 6. (a) Side-view (elevation) and (b) top-view (plan) of the electron density distribution of absorbed CO2 on Vs-ReS2. (c) Schematic of photocatalytic CO2 reduction in CR12 system under visible-light illumination (λ ≥ 420 nm). The purple-, red-, yellow-, orange-, white-, and black-colored spheres denote Cd, O, S, Re, H, and C atoms, respectively [51]. Reproduced with permission: Copyright 2021, Wiley-VCH Verlag GmbH & Co. KGaA. (d) Schematic representation of Au-Pt/Cu2O/ReS2 synthesis via the approaches of chemical reduction and photodeposition. (e) HRTEM images of Au-Pt/Cu2O/ReS2. Selectivity of CH4 (f) and CO (g) of Au-Pt/Cu2O/ReS2 with different Au-Pt mass ratios. (h) Schematic diagram of the transfer processes of the photoexcited electrons and holes and the mechanism of the photocatalytic CO2 reduction [111]. Copyright 2022, ELSEVIER.
Figure 6. (a) Side-view (elevation) and (b) top-view (plan) of the electron density distribution of absorbed CO2 on Vs-ReS2. (c) Schematic of photocatalytic CO2 reduction in CR12 system under visible-light illumination (λ ≥ 420 nm). The purple-, red-, yellow-, orange-, white-, and black-colored spheres denote Cd, O, S, Re, H, and C atoms, respectively [51]. Reproduced with permission: Copyright 2021, Wiley-VCH Verlag GmbH & Co. KGaA. (d) Schematic representation of Au-Pt/Cu2O/ReS2 synthesis via the approaches of chemical reduction and photodeposition. (e) HRTEM images of Au-Pt/Cu2O/ReS2. Selectivity of CH4 (f) and CO (g) of Au-Pt/Cu2O/ReS2 with different Au-Pt mass ratios. (h) Schematic diagram of the transfer processes of the photoexcited electrons and holes and the mechanism of the photocatalytic CO2 reduction [111]. Copyright 2022, ELSEVIER.
Molecules 28 02627 g006
Figure 7. (a) Schematic illustration of BaTiO3@ReS2 system preparation process. (b) TEM image of BaTiO3@ReS2. (c) Comparison of the piezo-catalytic, photocatalytic, and piezo-assisted photocatalytic degradation of RhB in the presence of BaTiO3@ReS2. (d) Schematic illustration of the synergy of photocatalysis and piezotronics effects on piezo-assisted photocatalysis system [115]. Reproduced with permission: Copyright 2022, Wiley-VCH Verlag GmbH & Co. KGaA.
Figure 7. (a) Schematic illustration of BaTiO3@ReS2 system preparation process. (b) TEM image of BaTiO3@ReS2. (c) Comparison of the piezo-catalytic, photocatalytic, and piezo-assisted photocatalytic degradation of RhB in the presence of BaTiO3@ReS2. (d) Schematic illustration of the synergy of photocatalysis and piezotronics effects on piezo-assisted photocatalysis system [115]. Reproduced with permission: Copyright 2022, Wiley-VCH Verlag GmbH & Co. KGaA.
Molecules 28 02627 g007
Figure 8. (a) Schematic illustration of the preparation process. (b) The HR-TEM image of rCQDs/ReS2 heterojunctions. Inset: The corresponding TEM image. (c) Reduction curves of the Cr (VI) aqueous solutions containing different photocatalysts under simulated sunlight irradiation: no catalyst, ReS2 nanosheets, CQDs/ReS2 heterojunctions, and rCQDs/ReS2 heterojunctions. (d) The charge density difference of rCQD(CQD)/ReS2 heterojunctions. (e) Schematic presentation of the carrier transfer process in the rCQD(CQD)/ReS2 heterostructure during reduction [86]. Reproduced with permission: Copyright 2022, ELSEVIER.
Figure 8. (a) Schematic illustration of the preparation process. (b) The HR-TEM image of rCQDs/ReS2 heterojunctions. Inset: The corresponding TEM image. (c) Reduction curves of the Cr (VI) aqueous solutions containing different photocatalysts under simulated sunlight irradiation: no catalyst, ReS2 nanosheets, CQDs/ReS2 heterojunctions, and rCQDs/ReS2 heterojunctions. (d) The charge density difference of rCQD(CQD)/ReS2 heterojunctions. (e) Schematic presentation of the carrier transfer process in the rCQD(CQD)/ReS2 heterostructure during reduction [86]. Reproduced with permission: Copyright 2022, ELSEVIER.
Molecules 28 02627 g008
Figure 9. Photocatalytic sterilization efficiency to Escherichia coli of different Ag-content samples grown on porous carbon cloth substrate, (a) under dark condition, (b) under visible light. (c) Schematic illustration of charge separation and photocatalytic disinfection via reactive oxygen species (ROS). (d) Photos of Escherichia coli colonies corresponding to photocatalytic sterilization for 0 min, 10 min, 20 min, and 30 min [85]. Reproduced with permission: Copyright 2022, ELSEVIER.
Figure 9. Photocatalytic sterilization efficiency to Escherichia coli of different Ag-content samples grown on porous carbon cloth substrate, (a) under dark condition, (b) under visible light. (c) Schematic illustration of charge separation and photocatalytic disinfection via reactive oxygen species (ROS). (d) Photos of Escherichia coli colonies corresponding to photocatalytic sterilization for 0 min, 10 min, 20 min, and 30 min [85]. Reproduced with permission: Copyright 2022, ELSEVIER.
Molecules 28 02627 g009
Table 1. Applications of ReS2-based heterostructure for hydrogen production.
Table 1. Applications of ReS2-based heterostructure for hydrogen production.
PhotocatalystSource of LightMorphologyPerformanceRef
ReS2/ZnIn2S4300 W Xenon arc lamp (λ ≥ 420 nm cutoff filter)cobblestone structure with particles1858.6 μmol h−1 g−1[61]
ReS2 nanowalls300 W Xenon arc lamp (λ ≥ 420 nm cutoff filter)ReS2 nanowalls13 mmol h−1 g−1[33]
TiO2/ReS2300 W Xenon arc lampReS2 nanosheets and TiO2 nanofibers1404 μmol h−1 g−1[58]
ReS2/TiO2Solar simulator (λ ≥ 300 nm)sea-urchin-like structured3.71 mmol h−1 g−1[104]
ReS2-BzO-TiO2Solar simulatornanosheets9.5 mmol h−1 g−1[56]
CdS/ReS2300 W Xennon lamp (λ ≥ 420 nm UV-cutoff filter)nanorod137.5 mmol h−1 g−1[59]
ReS2/Zn0.5Cd0.5Svisible light irradiation (λ ≥ 420 nm)nanospheres on ReS2 nanosheets112.10 mmol h−1 g−1[105]
ReS2/Mn0.2Cd0.8S300 W Xenon arc lamp (λ ≥ 420 nm cutoff filter)cauliflower-like morphology17.31 mmol h−1 g−1[43]
CdS/(Au-ReS2)300 W Xenon arc lamp (λ ≥ 420 nm cutoff filter)ReS2 nanosheets3060 μmol h−1 g−1[45]
g-C3N4/CdS/ReS2300 W Xenon arc lamp (AM 1.5 G filter)a hollow spherical nano-shell structure7141.2 ± 85.7 μmol h−1 g−1[44]
ReS2/ZnIn2S4-Sv300 W Xenon arc lampnanoflower1.08 mmol h−1 g−1[93]
ReS2/TiO2300 W Xenon arc lampReS2 ultrathin nanosheets1037 μmol h−1 g−1[34]
ReS2/g-C3N4Solar simulator (AM 1.5G)nanospheres1823 mmol h−1 g−1[92]
ReS2/ZnIn2S4Xenon arc lamp (400 nm cutoff light filter)ReS2 nanosheets2515 µmol h−1 g−1[50]
ReS2/g-C3N4300 W Xenon arc lampultrathin layered 2D/2D structure3.46 mmol h−1 g−1[54]
MoS2/ReS2@CdS300 W Xenon arc lamp (λ ≥ 420 nm)CdS@ReS2 nano-spheres and MoS2 nanoflakes171.9 mmol h−1 g−1[63]
ReS2/TiO2300 W Xenon arc lampcircle-shaped sheet-like structures 2D TiO2762.3 mmol h−1 g−1[103]
ReS2/g-C3N4visible light irradiation (λ ≥ 420 nm)ReS2 nanoflowers on the surface of g-C3N4249 μmol h−1 g−1[46]
ReS2-CdS/P-0.2300 W Xenon arc lamp (λ ≥ 420 nm cutoff filter)CdS nanorods and ReS2 nanosheet14.68 mmol h−1 g−1[62]
Ta3N5/ReS2300 W Xenon arc lampCdS nanorods and ReS2 nanosheet615 μmol h−1 g−1[53]
Table 2. Applications of ReS2-based heterostructure for CO2 reduction.
Table 2. Applications of ReS2-based heterostructure for CO2 reduction.
PhotocatalystSource of LightMorphologyPerformanceRef
ReS2@Cu2O/Cu300 W Xennon lamp (λ ≥ 420 nm UV-cutoff filter)ReS2 particles on the surface of Cu2O/Cu frameworksCO (14.3 μmol h−1 g−1)[64]
Au-Pt/Cu2O/ReS2300 W Xennon lampflower-like microsphereCH4 (60.76 μmol h−1 g−1)[111]
ReS2/CdSvisible-light irradiation (λ ≥ 420 nm)nanosheetsCO (7.1 μmol h−1 g−1)[51]
Table 3. Applications of ReS2-based photocatalysts of pollutants removal.
Table 3. Applications of ReS2-based photocatalysts of pollutants removal.
PhotocatalystTypeSynthesis MethodsMorphologyLight SourceApplicationEfficiencyCycleRef
ReS2/MIL-88B(Fe)Type-IIsolvothermal methodshuttle structureBoth PS and visible light irradiationDegradation of Ibuprofen (IBP)100% (3 h)3[52]
TiO2@ReS2Z-schemeultrasonic liquid exfoliation methodReS2 nanosheetsSolar simulatorDegradation of Rhodamine B (RhB)94% (120 min)25[47]
BaTiO3@ReS2Type-Imulti-step hydrothermal methodReS2 nanosheets on BaTiO3 nanorodsUV-vis lightDegradation of RhB96% (25 min)3 [115]
carbon quantum dots (CQDs)/ReS2Type-Itwo-step hydrothermal methodrCQDs/ReS2 nanosheets300 W Xenon lampCr (VI) 96% (50 min)6[86]
ReS2/Graphite Carbon Nitride (CN)Type-IIelectrostatic assembly processReS2 microspheres and CN nanosheets300 W Xenon arc lamp (λ ≥ 420 nm cutoff filter)RhB94% (30 min)3[116]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, N.; Li, Y.; Wang, L.; Yu, X. Photocatalytic Applications of ReS2-Based Heterostructures. Molecules 2023, 28, 2627. https://doi.org/10.3390/molecules28062627

AMA Style

Wang N, Li Y, Wang L, Yu X. Photocatalytic Applications of ReS2-Based Heterostructures. Molecules. 2023; 28(6):2627. https://doi.org/10.3390/molecules28062627

Chicago/Turabian Style

Wang, Nan, Yashu Li, Lin Wang, and Xuelian Yu. 2023. "Photocatalytic Applications of ReS2-Based Heterostructures" Molecules 28, no. 6: 2627. https://doi.org/10.3390/molecules28062627

APA Style

Wang, N., Li, Y., Wang, L., & Yu, X. (2023). Photocatalytic Applications of ReS2-Based Heterostructures. Molecules, 28(6), 2627. https://doi.org/10.3390/molecules28062627

Article Metrics

Back to TopTop