Next Article in Journal
Ni-Based Catalysts: Synthesis and Applications Today
Next Article in Special Issue
Synthesis of g-C3N4@ZnIn2S4 Heterostructures with Extremely High Photocatalytic Hydrogen Production and Reusability
Previous Article in Journal
Nd2−xSrxNiO4 Solid Solutions: Synthesis, Structure and Enhanced Catalytic Properties of Their Reduction Products in the Dry Reforming of Methane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress of ZnIn2S4-Based Catalysts for Photocatalytic Overall Water Splitting

1
School of Environmental and Chemical Engineering, Jiangsu University of Science and Technology, Zhenjiang 212003, China
2
School of Material Science and Engineering, Jiangsu University of Science and Technology, Zhenjiang 212003, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(6), 967; https://doi.org/10.3390/catal13060967
Submission received: 10 May 2023 / Revised: 30 May 2023 / Accepted: 31 May 2023 / Published: 2 June 2023

Abstract

:
Photocatalytic overall water splitting in solar–chemical energy conversion can effectively mitigate environmental pollution and resource depletion. Stable ternary metal indium zinc sulfide (ZnIn2S4) is considered one of the ideal materials for photocatalytic overall water splitting due to its unique electronic and optical properties, as well as suitable conduction and valence band positions for suitable photocatalytic overall water splitting, and it has attracted widespread researcher interest. Herein, we first briefly describe the mechanism of photocatalytic overall water splitting, and then introduce the properties of ZnIn2S4 including crystal structure, energy band structure, as well as the main synthetic methods and morphology. Subsequently, we systematically summarize the research progress of ZnIn2S4-based photocatalysts to achieve overall water splitting through modification methods such as defect engineering, heterostructure construction, and co-catalyst loading. Finally, we provide insights into the prospects and challenges for the overall water splitting of ZnIn2S4-based photocatalysts.

1. Introduction

The world is currently suffering from environmental pollution and resource depletion, with energy issues looming large. According to relevant studies, the annual global consumption of energy is equivalent to the solar energy reaching the Earth’s surface every hour; therefore, solar energy as an abundant, non-polluting natural resource has replaced the traditional fuel fossil as a research hotspot [1]. However, solar energy has limitations such as intermittency and low density, so an effective storage method is needed to make efficient use of solar energy [2]. Since 1972, when it was reported that TiO2 semiconductors could produce hydrogen and oxygen when irradiated by ultraviolet light, photocatalysis, which uses solar energy to convert it into storable chemical energy, has attracted extensive research [3].
Hydrogen, as a clean, high-energy-density solar fuel, is the ideal energy carrier. Since most photocatalytic hydrogen production studies require the use of sacrificial agents to achieve this, photocatalytic overall water splitting is considered a low-cost, ideal method for converting solar energy into hydrogen energy [4,5,6]. The photocatalytic overall water splitting process is based on three fundamental photocatalytic processes: photocatalyst absorption of photons to generate electron–hole pairs, photogenerated charge transfer and separation, and surface redox reactions. A variety of semiconductor catalysts such as metal oxides, metal sulfides, and nitrides are currently used in the field of photocatalytic overall water splitting [7,8,9,10,11,12]. Among them, metal sulfides have the advantages of good charge transfer ability, suitable energy band structure for overall water splitting, and excellent light collection ability to become one of the potential catalysts in photocatalytic overall water splitting [13].
Metal sulfides are mainly classified into binary metal sulfides such as CdS, MoS2, and ZnS; ternary metal sulfides such as ZnIn2S4 and CuInS4; and polymetallic sulfides such as AgZnInS [14]. Most of these binary sulfides have some disadvantages that are more difficult to improve, such as ZnS-based photocatalysts having a poor photo-response, responding only to ultraviolet (UV) light, and CdS-based catalysts having severe photo-corrosion and poor stability, whereas ternary metal sulfides tend to be more stable [15,16,17,18]. Zinc indium sulfide (ZnIn2S4), a ternary metal sulfide belonging to the AB2X4 family, has unique electronic and optical properties. Compared with conventional photocatalysts, ZnIn2S4 has a narrower band gap, adjustable between about 2.06 and 2.85 eV, and has thermodynamically suitable conduction and valence band positions for photocatalytic overall water splitting as well as a strong visible-light response range [19,20]. In addition, ZnIn2S4 has many advantages such as strong photostability, relatively environmentally friendly chemical composition, ease of preparation, and wide distribution of raw materials [21]. Therefore, ZnIn2S4 is a more desirable material for photocatalytic overall water splitting.
Although ZnIn2S4 has many advantages, in practical applications, ZnIn2S4-based photocatalysts suffer from difficulties in achieving one-component photocatalytic overall water splitting or low photocatalytic overall water splitting efficiency, mainly due to the slow photo-generated charge separation and migration efficiency and weak solar energy utilization [22,23,24]. Therefore, appropriate modification strategies such as elemental doping, vacancy engineering, the construction of heterojunctions, and the loading of co-catalysts are required to improve the performance of ZnIn2S4-based photocatalyst materials.
Researchers have actively explored how to improve the performance of ZnIn2S4-based photocatalysts and have reported on a review of ZnIn2S4 photocatalysts from different perspectives. For example, Liu et al. reviewed the research progress of ZnIn2S4-based photocatalysts constructed with heterojunctions for photocatalytic hydrogen production [25]. Yadav et al. reviewed various modification strategies to improve the performance of ZnIn2S4-based photocatalysts and summarized their applications in water pollution treatment, CO2 reduction, etc. [26]. However, previous reports are mainly based on applications such as hydrogen production and pollutant treatment, and there is no systematic summary of the research progress on ZnIn2S4-based photocatalysts for achieving photocatalytic overall water splitting.
Hence, this paper systematically reviews the research progress of ZnIn2S4-based photocatalysts in photocatalytic overall water splitting. First, we briefly describe the mechanism of photocatalytic overall water splitting. Then, we outline the properties of ZnIn2S4, including its crystal structure, energy band structure, main synthesis methods, and morphology. The modification strategies of ZnIn2S4-based photocatalysts are reviewed, mainly including surface engineering such as doping and vacancies, the construction of heterojunctions, and the loading of co-catalysts (Scheme 1). Finally, we provide an outlook on the prospects and challenges of ZnIn2S4-based photocatalysts for photocatalytic overall water splitting. There are almost no reviews on ZnIn2S4-based photocatalytic overall water splitting, and the study in this paper provides the latest research progress on ZnIn2S4-based water splitting catalysts, which is important for the design and synthesis of efficient ZnIn2S4-based photocatalytic overall water splitting catalysts.

2. Mechanisms of Photocatalytic Overall Water Splitting

Photocatalytic overall water splitting consists of two half-reactions, the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER). Theoretically, in order to achieve overall water splitting, the semiconductor band gap should be no less than 1.23 eV under standard conditions, the potential at the bottom of the conduction band should be less than 0 eV (H2/H+ = 0 eV vs. NHE, pH = 0), and that at the top of the valence band should be greater than 1.23 eV (H2O/O2 = 1.23 eV vs. NHE, pH = 0) [27]. The redox potential of water is all located within the band gap of the photocatalyst and photocatalytic overall water splitting is thermodynamically feasible. However, photocatalytic overall water splitting is an uphill reaction requiring additional energy to promote water splitting, which is a thermodynamically unfavorable process (G > 0); therefore, hydrogen and oxygen are prone to the reverse reaction and H2O reformation, which severely inhibits the photocatalytic water splitting activity [28].
Semiconductor-based photocatalysts for photocatalytic overall water splitting are based on three basic processes of photocatalysis: under solar irradiation with an energy greater than the band gap of the photocatalyst, photogenerated electrons are excited to leap to the conduction band and photogenerated holes remain in the valence band; photogenerated charges migrate separately to the semiconductor reaction site; and un-recombined photogenerated electrons and holes undergo redox reactions of water at the catalyst surface [29]. From the kinetic point of view, the recombination of photogenerated carriers is much faster than their redox reactions at the surface. The Coulomb force constraints between photogenerated charges and high interfacial potential barriers during charge transfer lead to rapid photogenerated carrier recombination and low utilization efficiency, which severely limit photocatalytic activity [30].
In addition, the range of solar energy utilization affects the photocatalytic activity. According to relevant research reports, the UV content of natural sunlight is less than 3%, the visible content is less than 40%, and the near-infrared occupies about 50% of the sunlight, while photocatalytic materials capture light basically in the UV and visible region, with a low efficiency of solar energy utilization [31,32]. The overall photocatalytic water splitting activity is limited by the low light collection capacity of the catalyst, the rate of photogenerated charge separation and migration, and the surface oxidation reaction [33]. Therefore, researchers have adopted corresponding modification strategies to prepare photocatalysts with high activity and high solar energy utilization efficiency. The stable ternary metal sulfide ZnIn2S4 is one of the ideal materials for photocatalytic overall water splitting due to its advantages. As shown in Scheme 2, the ZnIn2S4-based photocatalysts have thermodynamically suitable conduction and valence band positions for photocatalytic water splitting. However, single-component photocatalytic water splitting is difficult to achieve due to the overall low charge utilization and solar utilization as well as photo-corrosion phenomena. Therefore, modification strategies such as the doping of heteroatoms, formation of defects, construction of heterojunctions, and loading of co-catalysis were adopted to enhance the ZnIn2S4-based photocatalytic performance.

3. Introduction of ZnIn2S4

3.1. Crystal and Energy Band Structures

The ternary metal sulfide ZnIn2S4 has three crystal phase structures: cubic, hexagonal, and rhombic. As shown in Figure 1a, the cubic ZnIn2S4 has an ABC stack in which the Zn atom is tetrahedrally coordinated to the S atom and the In atom is octahedrally coordinated to the S atom [34]. As shown in Figure 1b, the hexagonal ZnIn2S4 has a layered structure in which the atoms are repeatedly stacked in the order S-Zn-S-In-S-In-S. In the hexagonal crystal phase, the Zn atoms and half of the In atoms form a tetrahedral coordination with the S atoms, and the remaining In atoms form an octahedral coordination with the S atoms [35]. As shown in Figure 1c, the crystal structure of rhombic ZnIn2S4 is similar to that of hexagonal ZnIn2S4, consisting of a sandwich layer with one octahedron and two tetrahedra. Rhombic ZnIn2S4 differs from hexagonal ZnIn2S4 in that the Zn atoms and half of the In atoms are mixed in the tetrahedral sites [36]. Different crystalline phases have different properties, the cubic phase has thermoelectric properties, the hexagonal phase has photoconductivity, and the rhombic phase has good charge transfer ability. In addition, the researchers found that the photocatalytic activity of ZnIn2S4 can be effectively enhanced by changing the crystalline phase. For example, in 2011, Shen et al. synthesized ZnIn2S4 with different crystalline phase structures by high-temperature thermal sulfide treatment of metal oxide precursors [37]. They performed thermal sulfidation reactions on Zn-In mixed oxide precursors under an H2S atmosphere and synthesized cubic ZnIn2S4 and gradually transformed it into rhombohedra when the thermal sulfidation temperature was increased from 400 to 800 °C. The rhombohedral ZnIn2S4 has a good charge transfer ability and light absorption, enhancing the photocatalytic hydrogen production activity under sacrificial agents. In addition, ZnIn2S4 can be used to transform the crystalline phase by using different metal precursors and changing the reaction temperature to prepare more active crystalline phases and structures for photocatalytic applications. Density functional theory (DFT) has been widely used in semiconductor materials to study the electronic structure of materials. DFT provides theoretical insight into the electronic energy band structure of cubic, hexagonal, and rhombic ZnIn2S4, and all three crystalline phases are direct band gap semiconductors. Although the respective band gaps of the cubic, hexagonal, and rhombic crystalline phases are known from theoretical calculations, the actual band gaps deviate from the calculated results due to the limitations of the local density approximation (LDA) functional [38].

3.2. Synthetic Methods and Morphology of ZnIn2S4

3.2.1. Morphology of ZnIn2S4

The conformation and structure of ZnIn2S4, a stable ternary metal sulfide semiconductor, significantly affect the photocatalytic activity. Among them, ZnIn2S4 with different morphological structures such as zero-dimensional (0D), one-dimensional (1D), two-dimensional (2D), and three-dimensional (3D) plays an important role in the solar–chemical energy conversion process.
Zero-dimensional ZnIn2S4 is mainly available in quantum dots (QD). Quantum dots, as a hot research material in the energy environment, have the advantages of easy synthesis, abundant surface sites, and controllable size, and due to the quantum confinement effect, their light collection range can be adjusted to the near-infrared region, which significantly improves the solar light utilization effect [39]. Peng et al. prepared size-tunable ZnIn2S4 quantum dots using oleylamine as the ligand and uncoordinated octadecene as the solvent [40]. Due to the different effects of reaction temperature in controlling the nucleation and growth process of nanocrystals thermodynamically and kinetically, ZnIn2S4 nanocrystals with sizes ranging from 2.1 to 10.1 nm were synthesized by varying the reaction temperature between 140 °C and 210 °C. Experiments have shown that ZnIn2S4 nanocrystals with small size and annealed to remove the capping agent are highly active in the degradation of methylene orange. Currently, 1D ZnIn2S4 is mainly available as nanowires, nanotubes, and nanoribbons, but most 1D ZnIn2S4 synthesis requires the use of templates, so most of the synthesized morphologies are 2D and 3D ZnIn2S4 [41,42]. The majority of ZnIn2S4 syntheses are hexagonal due to the advantages of good stability, simple preparation, and high activity of the hexagonal crystalline phase ZnIn2S4. It is easy for the hexagonal ZnIn2S4 layered structure to form a 2D nanosheet morphology; especially, the ultra-thin 2D nanosheets have the advantages of short photogenerated carrier migration distance, large specific surface area, and abundant surface-active sites, which significantly improve the photocatalytic activity of ZnIn2S4-based materials [43]. In 2018, Zhang et al. prepared ultrathin ZnIn2S4 nanosheets (Vs-M-ZnIn2S4) rich in S-vacancies by exfoliating large blocks of ZnIn2S4 synthesized using lithium intercalation [44]. Then, MoS2 quantum dots (MoS2QDs) were grown in the S-vacancy region of Vs-M-ZnIn2S4 induced by S-vacancies in one of the Zn cuts to synthesize atomic-level heterojunction MoS2QDs@Vs-M-ZnIn2S4. Unlike ZnIn2S4, the S-vacancies in MoS2QDs@Vs-M-ZnIn2S4 act as electron traps, enriching electrons in the Zn plane and transferring them to MoS2QDs via Zn-S bonds, preventing the rapid recombination of photogenerated charges due to vertical electron transport and improving the hydrogen production activity of the photocatalyst.
However, due to the high surface energy of the individual nanosheets, ZnIn2S4 tends to aggregate to form 3D microspheres during growth, and both the specific configuration of its synthesis and the increased specific surface area and active sites enhance the photocatalytic activity of the catalyst. Shen et al. synthesized persimmon-layered ZnIn2S4 photocatalysts by using an oleylamine (OA)-assisted solvent method, adding tetrahydrofuran (THF) solution to form a hexagonal structure and OA selectively adsorbed on the ZnIn2S4 hexagon to form nanoplates, which then self-assembled to form persimmon [45]. The material has excellent photocatalytic hydrogen production activity, with the best catalyst achieving 220.45 mmol h−1 after 3% Pt loading. Some of the ZnIn2S4 photocatalysts synthesized in current research, such as marigold and peony-flower-like, as well as rose and marigold-like, are 3D microspheres [46,47]. Furthermore, hollow structured materials with their rich surface-active sites, short charge migration paths, and good light collection capabilities have attracted the interest of a wide range of researchers [48]. In 2014, Warule et al. synthesized hollow marigold-like ZnIn2S4 materials for photocatalytic hydrogen production by a surfactant-assisted hydrothermal method [49]. The hollow marigold-like nanoparticles have a higher specific surface area and more active sites, exhibiting better hydrogen production activity. As most photocatalytic reactions occur on the surface of the material, the morphology and configuration of the material have an important influence on the photocatalytic activity. The synthesis of materials with high specific surfaces, abundant active sites, and unique configurations is very important for improving the performance of ZnIn2S4. We provide a brief summary of the various morphologies of ZnIn2S4 in Table 1.

3.2.2. Synthetic Methods of ZnIn2S4

Hydrothermal method. Unlike other synthesis methods that use templates and special equipment, the hydrothermal synthesis of ZnIn2S4 is relatively gentle and simple. The reaction temperature, the type of reaction precursor, the pH of the reaction system, the use of surfactants, and the type of surfactant all influence the synthesis, morphology, and crystallinity of the ZnIn2S4-based catalysts. In 2006, Guo et al. prepared ZnIn2S4 solid and hollow microsphere structures hydrothermally under cetyltrimethylammonium bromide (CTAB) and ethylene glycol (PEG) as surfactants, respectively [52]. Two different ZnIn2S4 crystalline phase materials were hydrothermally synthesized by the addition of different metal precursors by Chen et al. in 2012 [53]. The cubic phase was synthesized when metal nitrates were added as precursors, whereas a thermodynamically stable hexagonal phase was prepared when metal chlorides containing electronegative low-chloride ions were used as precursors. In addition, Warule et al. in 2014 synthesized hollow marigold-like and rose-shaped ZnIn2S4 photocatalysts by a surfactant-assisted hydrothermal method using polyvinylpyrrolidone (PVP) and diethylamine (DEA), respectively [49]. Studies have shown that varying the concentration of PVP (100 ppm–300 ppm) results in the synthesis of different forms ranging from twisted to hollow marigold-like. In 2022, Yin et al. synthesized ZnIn2S4 with different assembled microstructures for dehydrogenation treatment by a one-step hydrothermal method and investigated the effect of different solvents on the structure of ZnIn2S4 [54]. The experimental results showed that the solvents were water, ethanol, and ethylene glycol in accordance with which smooth surface-petal-like, relatively rough micro-disk-like and thin nanosheets were synthesized. In 2022, Zou et al. hydrothermally prepared ZnIn2S4 materials with different crystalline phases by adjusting the pH of the system using oxalic acid as a chelating agent [55]. The cubic phase of ZnIn2S4 was prepared without the addition of oxalic acid and hydrothermal preparation after oxalic acid modification transformed the original cubic phase of ZnIn2S4 into a hexagonal phase.
Solvothermal method. Like hydrothermal methods, solvothermal synthesis is relatively simple and is widely used in the synthesis of materials for the energy environment. The nature of the solvent such as alkalinity, viscosity, and type of solvent play an important role in the preparation of ZnIn2S4 materials in terms of their morphology. Guo et al. in 2006 found that reaction temperature has a significant effect on ZnIn2S4 morphology [52]. They prepared the catalysts by varying the reaction temperature solvothermally using pyridine as the solvent; when the reaction temperature was between 120 and 160 °C, ZnIn2S4 grew in the (002) direction to form nanoribbons, and when the reaction temperature was 180 °C and higher, the material formed nanotubular shapes. In addition to the reaction temperature, the type of solvent has an equally important influence on the synthetic microstructure of ZnIn2S4. In 2008, Shen et al. used water, methanol, and ethylene glycol as solvents, with the first two synthesizing cherry-shaped microspheres and ethylene glycol as the solvent to synthesize micro-clustered clusters [56]. The results show that the catalysts synthesized with water as the solvent are more crystalline and stable and have the best hydrogen production efficiency. Furthermore, Su et al. in 2016 used water, ethanol, methanol, as well as ethylene glycol as solvent precursors for the synthesis of ZnIn2S4 by a solvothermal method to investigate the effect of different types of solvents on the selective oxidation activity of aromatic alcohols [57]. The experimental structures showed that the best photocatalytic performance of ZnIn2S4 was prepared using ethanol as the solvent, and that the difference in performance between the different solvents was mainly due to the degree of exposure of the basic crystalline surface, resulting in different exposures of the special surface.
Other synthesis methods. In addition to hydrothermal and solvothermal methods, microwave-assisted methods, thermal sulfur methods, chemical vapor deposition, and spray pyrolysis are also applied to the synthesis of ZnIn2S4 in different forms and structures [58,59,60]. In 2012, Huang et al. synthesized ZnIn2S4 in micro-spherical form by a simple spray pyrolysis method [61]. They first synthesized the solution by magnetic stirring, then atomized the solution and fed it into a tubular reactor for reaction to produce spherical particles. In addition, Pop et al. in 2022 prepared ZnIn2S4 photocatalysts by the microwave-assisted oil bath production method to investigate the effect of ZnIn2S4 materials with different zinc concentrations and reaction temperatures on methyl orange adsorption [62]. The experimental results show that disordered cubic ZnIn2S4 is synthesized at 160 °C and hexagonal phase ZnIn2S4 structures are synthesized at 180 °C. Moreover, the morphology changes from nanoparticles to nanoflower, hollow-microsphere structures as the zinc concentration and reaction temperature increase. Although there are numerous methods to synthesize ZnIn2S4, hydrothermal and solvothermal methods are widely used in the synthesis of ZnIn2S4-based photocatalysts due to the advantages of the simple synthesis process and low cost.
Since most catalytic reactions occur on the surface of catalysts, the structure and morphology of materials have important effects on their properties. Therefore, we briefly describe representative morphologies in ZnIn2S4 materials as well as hydrothermal and solvothermal synthesis methods. It is hoped that this will provide some reference for the preparation of ZnIn2S4 photocatalytic overall water splitting morphologies using simple synthetic methods.

4. Photocatalytic Overall Water Splitting Modification Based on ZnIn2S4 Catalyst

4.1. Defect Engineering

Defect engineering not only modulates the electronic structure, provides abundant active sites, and improves the efficiency of charge separation and migration of materials, but also changes the morphology and interfacial reactions of materials and plays an important role in the catalytic reactions of materials [63]. Therefore, defect engineering is often used in photocatalytic systems to modulate material properties, among which metal sulfides have received a lot of attention because of their tunable electronic structure. Doping and vacancies, which are point defects in defect engineering, have been applied to modulate material properties in sulfide catalysts such as ZnIn2S4. Su et al. synthesized Al3+-doped strontium calcium titanate photocatalysts (Al-STO) for photocatalytic overall water splitting by polymerization complexation [64]. The doping of low-valence metal cations effectively promotes the migration rate of photogenerated carriers, and the appropriate oxygen vacancies on the surface facilitate the adsorption of water molecules and hydroxyl groups to promote the reduction reaction. The reduction in intrinsic Ti3+ defects effectively promotes the separation and migration efficiency of photogenerated carriers in concert with the oxygen vacancies on the surface to improve the photocatalytic activity of strontium-titanate-based photocatalysts. The best material—2% Al-STO—achieved hydrogen yields of 1.256 mmol h−1 and oxygen yields of 0.692 mmol h−1 under a loaded co-catalyst, with an apparent quantum efficiency (AQE) of up to 55.46% at 365 nm for the composite. It follows that the introduction of defects in semiconductor photocatalysts is a promising method of modification.

4.1.1. Doping Strategy

Doping strategies are modification methods that introduce impurity atoms, such as metals and non-metals, into the structure of photocatalytic materials to form defects to alter the material properties, mainly through both doping and ion exchange [65,66]. Material doping with heteroatoms has the advantages of narrowing the band gap of the composite, enhancing light absorption, changing the morphological structure of the material, and increasing the separation and migration rates of photogenerated carriers, and is therefore considered to be an effective method for effectively increasing the activity of sulfide photocatalysts [67]. Alkaline earth metals such as Ca and Ba, rare-earth metals such as La and Y, transition metals such as Ag, Co, and Ni, as well as non-metals such as N and O have been employed in ZnIn2S4, and good photocatalytic activity has been achieved [68].

Metal Doping

Pan et al. in 2021 reported Ag-ZnIn2S4 composite photocatalytic materials with dual defects in Ag doping and nanopores prepared by cation exchange between 2D ZnIn2S4 monolayers and Ag, achieving ZnIn2S4 photocatalytic overall water splitting without co-catalysts and sacrificial agents under visible light irradiation [69]. The experimental results show that Ag doping can effectively narrow the band gap, improve the light collection capacity of the catalyst, and promote photogenerated charge separation. The hydrogen yield of the composites was 56.6 μmol g−1 h−1 and the oxygen yield was 29.1 μmol g−1 h−1. The Ag-ZnIn2S4 composite has good photocatalytic activity, as shown in Figure 2a; the Ag element doping redistributes the charge; the positively charged enriched Ag adsorbs water molecules and promotes the oxygen production reaction; and the sulfur atoms suspended on the nanopores promote the hydrogen production reaction. In addition, the ultrathin 2D ZnIn2S4 monolayer narrows the migration distance and increases the photogenerated charge separation and migration rate. They also prepared Cu-ZnIn2S4 with a similar structure to Ag-ZnIn2S4 by the same method. The experimental results show that Cu doping has a stronger charge migration rate and higher hydrogen production activity, but as shown in Figure 2b, Cu doping does not enable photocatalytic all-water decomposition, which confirms that Ag doping is important for the Ag-ZnIn2S4 materials to achieve photocatalytic all-water decomposition.
In addition, in 2022, Sun et al. used a magnetic stirring one-step solvothermal method to synthesize in situ interstitial zinc-doped ultrathin ZnIn2S4 nanosheets (dZni-ZnIn2S4) and to study the photocatalytic performance of the materials for photocatalytic overall water splitting without any auxiliary agent under visible-light irradiation [70]. The Zn doping not only produces a static potential difference, but also forms a short range of disordered structures with abundant active sites. As shown in Figure 2e, the Zn doping leads to broadening and weakening of the ZnIn2S4 (006) peak and disruption of interlayer stacking, and widens the material nanosheet spacing, forming a short-range disordered structure. The differential charge density map (Figure 2c,d) reveals that, unlike the original ZnIn2S4 with uniform charge density distribution, dZni-ZnIn2S4 leads to charge redistribution with positive charge enrichment at the Zn site, lowering the water oxidation potential barrier and increasing the electron density at the sulfur site, which facilitates H+ adsorption reduction and improves the photocatalytic overall water splitting activity. In addition, the ultrathin nanosheets formed by the dZni-ZnIn2S4 material shorten the migration distance and effectively promote photogenerated carrier separation, achieving a hydrogen yield of 42.8 μmol g−1 h−1 and an oxygen yield of 19.1 μmol g−1 h−1.
Subsequently, Sun et al. in 2022 also used a magnetic stirring one-step solvothermal method to synthesize Al-ZnIn2S4 composites and investigate their hydrogen and oxygen production properties under visible-light irradiation [71]. Al-ZnIn2S4 has a rich mesoporous structure that maximizes the exposure of active sites and enhances photocatalytic activity. The Al doping also produces an expanded layer spacing that induces a domain electrostatic potential difference, effectively promoting photogenerated carrier separation. Al doping redistributes the charge, enriching the Al sites with positive charge, increasing the electron density around the S sites, promoting redox reactions in the material, reducing photo-corrosion, and improving stability. Due to these characteristics, the material achieves efficient photocatalytic overall water splitting without any auxiliary agent, with a hydrogen yield of 77.2 μmol g−1 h−1 and an oxygen yield of 35.3 μmol g−1 h−1.

Non-Metal Doping

Jing et al. prepared ZnIn2S4-350 °C-4 h with oxygen doping and S-vacancies by hydrothermal synthesis of hexagonal ZnIn2S4 by calcination at 350 °C under an air atmosphere for 4 h [72]. The calcined ZnIn2S4 improves light absorption, and the surface-rich S-vacancies extend the photogenerated charge lifetime to inhibit recombination but do not increase the carrier density. The increase in electron paramagnetic resonance (EPR) signal after calcination indicates that calcination introduces S-vacancies, which increases and then decreases with increasing calcination time, and that the surface oxygen doping gradually fills in. After air calcination, the Fourier-transform infrared (FTIR) spectrum shows SO42− bidentate characteristic peaks, further demonstrating the synergistic effect of the material between the oxygen doping and the S-vacancies. The synergistic effect of S-vacancies and oxygen doping on the material surface enhances photogenerated carrier lifetimes and concentrations and improves the photocatalytic overall water splitting activity. The materials are loaded with Pt/Cr co-catalyst and show remarkable photocatalytic activity with a hydrogen yield of 270.2 μmol g−1 h−1 and an oxygen yield of 130.0 μmol g−1 h−1, which are 13 times higher than those of the pristine ZnIn2S4.
In summary, compared with the undoped ZnIn2S4-based material, the material doping element can effectively reduce the band gap of the composite material, improve the material’s light collection capacity, change the morphological structure of the material, modulate the electronic structure of the material, and increase the photogenerated carrier separation and migration rate. Although the role of doping elements in photocatalysis remains in doubt due to their possible negative effects as photogenerated charge recombination centers, doping strategies do play an important role in modifying ZnIn2S4-based photocatalytic materials.

4.1.2. Vacancy Introduction

Vacancies are material lattice defects caused by the lack of cations or anions, which is mainly divided into cationic and anionic types [73]. He et al. prepared single-atom sulfur vacancy CdS nanorods (Sv-CdS NRs) with a spin-polarized electric field by the hydrothermal method combined with anaerobic heating to achieve photocatalytic overall water splitting without any co-catalyst [74]. Single-atom sulfur vacancies are introduced to induce spin-polarization properties in the CdS system, providing a strong facilitating electric field that enhances photogenerated carrier separation and migration rates. Sv-CdS NRs significantly improves the photocatalytic activity of the material, and the best material, Sv-CdS-2 NRs, was able to achieve a hydrogen evolution performance of 363.8 μmol g−1 h−1 and an oxygen evolution performance of 181.9 μmol g−1 h−1. In 2022, a rhombic ZnIn2S4-800 material rich in S-vacancies was reported by Jing et al. [75]. They designed and prepared a rhombic crystal phase and sulfur-vacancy-rich ZnIn2S4-800 material by calcining hexagonal ZnIn2S4 at high temperature under a nitrogen atmosphere, unlike the previous harsh preparation conditions, which changed the crystal phase of ZnIn2S4. As shown in Figure 3a, X-ray diffraction (XRD) patterns of ZnIn2S4 before and after calcination show that high-temperature calcination under a nitrogen atmosphere has modified the crystal phase and structural morphology of ZnIn2S4. Characterization by electron paramagnetic resonance (Figure 3b), UV-diffuse reflectance spectroscopy (Figure 3c), and the energy band structure (Figure 3d) confirms that the introduction of sulfur vacancies into ZnIn2S4 by calcination has narrowed the band gap and improved the light gathering capacity of the material. Appropriate S-vacancies as electron capture sites to increase the photogenerated carrier density inhibit recombination and promote hydrogen production reactions, and the calcined material has a smaller average mass of electrons and holes, accelerating the rate of photogenerated charge migration. The material loaded with Pt/Cr co-catalyst achieves hydrogen yields of 68.0 μmol g−1 h−1 and oxygen yields of 31.0 μmol g−1 h−1 in pure water under visible-light irradiation.
In summary, the introduction of appropriate vacancies in ZnIn2S4-based photocatalytic materials can narrow the band gap and improve visible-light absorption, increase the charge density of photogenerated carriers and the rate of migration of photogenerated charge separation, and regulate the electronic structure of the material, providing an abundance of surface-active sites and effectively modulating the crystal and energy band structure of the material, thus effectively improving the photocatalytic activity of the material.

4.2. Construction of Heterogeneous Junctions

The construction of heterojunctions from two or more semiconductors is a better modification strategy due to the strong Coulombic forces on the photogenerated charges of a single semiconductor, which are prone to recombination and result in poor photocatalytic activity. The construction of heterojunctions not only improves the photoresponse range of the material, provides a larger specific surface area, exposes more active sites, and modulates the catalyst to form unique structures, but also increases the photogenerated carrier separation and migration rates [76,77]. Depending on the energy band structure of the semiconductor and the photogenerated carrier transfer path, heterojunctions can be classified into six categories: type I heterojunction, type II heterojunction, Schottky junction, p–n heterojunction, Z-scheme heterojunction, and S-scheme heterojunction. The construction of a heterojunction by coupling ZnIn2S4 with another semiconductor is one of the effective modification methods to overcome the disadvantages of ZnIn2S4-based photocatalysts such as slow photogenerated charge separation and migration, and easy recombination [78,79]. Therefore, it is important to design and prepare heterojunctions with an appropriate energy band structure and well-matched geometry. In recent years, the photocatalytic performance of heterojunction materials constructed from metal oxides, carbon nitride, and ZnIn2S4 has been improved to varying degrees and applied in the field of environmental energy [80,81,82,83,84]. Currently, the most widely used in photocatalytic overall water splitting are Z-scheme heterojunctions and Schottky heterojunctions.

4.2.1. Z-Scheme Heterojunction

Most semiconductor materials are unable to achieve one-step photo-excited overall water splitting, due to drawbacks such as the lack of a suitable energy band structure and severe photogenerated carrier complexation. In recent years, the use of two-step photoexcitation, coupling two or more semiconductors to construct a Z-scheme heterojunction, has been widely used to improve the photocatalytic activity of ZnIn2S4-based photocatalysts with good results [85,86]. There are two main types of Z-scheme heterojunctions: (1) Indirect Z-scheme heterojunctions in which the two semiconductors are not in contact and the charge is transferred through a charge transfer medium such as Au and graphene; (2) a direct Z-scheme heterojunction in which two semiconductors come into contact, generating an electric field at the inner boundary due to differences in the work function and Fermi energy levels, accelerating the separation of photogenerated carriers [87,88]. The construction of Z-scheme heterojunctions facilitates the spatial separation of photogenerated charges, inhibits recombination, and maintains the high redox capacity of the material, effectively improving photocatalytic activity.

Indirect Z-Scheme Heterojunction

In 2017, Zhong et al. reported an indirect Z-scheme heterojunction material, ZnIn2S4/RGO/BMO, in which ZnIn2S4 acts as a hydrogen-depleting photocatalyst, Bi2MoO6 (BMO) acts as an oxygen-depleting photocatalyst, and reduced graphene oxide (RGO) acts as an electron mediator (Figure 4a) [89]. They prepared BMO and ZnIn2S4 by solvent thermal and hydrothermal methods, respectively, and then loaded Pt and CoOx co-catalysts onto ZnIn2S4 and BMO, respectively, to achieve photocatalytic overall water splitting without sacrificial agents. The construction of Z-scheme heterojunctions and the appropriate number of electron-mediated RGO can effectively increase the photogenerated charge separation and migration rates of the catalysts and improve the photocatalytic activity of the composites. Experiments show that the optimum material is Pt/ZnIn2S4-RGO (3%)-CoOx/BMO, which achieves a hydrogen yield of 31.4 μmol g−1 h−1 and an oxygen yield of 15.8 μmol g−1 h−1 under visible-light irradiation in pure water.
In 2018, Yang et al. reported a sea-urchin-shaped ZnIn2S4-based catalyst (ZnIn2S4-Au-TiO2), achieving a hydrogen yield of 186.3 μmol g−1 h−1 and an oxygen yield of 66.3 μmol g−1 h−1 without co-catalysts and sacrificial agents [90]. They first synthesized TiO2 microspheres, then synthesized Au-TiO2 by chemically depositing Au nanoparticles on the TiO2 surface, and finally prepared the indirect Z-scheme heterojunction material ZnIn2S4-Au-TiO2 by direct deposition of ZnIn2S4 on the Au-TiO2 surface via a solvothermal method (Figure 4b). The Z-scheme heterojunction can effectively promote the electron–hole pair separation efficiency, and the ultra-thin ZnIn2S4 nanosheets increase the specific surface area of the material and provide more active sites. As shown in Figure 4c, the ultraviolet–visible (UV-Vis) diffuse reflectance spectra show an increase in the visible absorption capacity of the material due to the surface plasmon resonance of the gold nanoparticles (Au NPs). The composite material promotes the separation of photogenerated carriers and increases the light collection capacity of the material, thus effectively improving the photocatalytic activity of the material.
In 2021, Geng et al. reported a Pt-ZnIn2S4/RGO/Co3O4-BiVO4 (110) photocatalyst [92]. They first prepared BiVO4 by the solvothermal method, RGO/Co3O4-BiVO4 (110) by photo-deposition and reduction, and Pt-ZnIn2S4 by the hydrothermal method and photo-deposition. Finally, they used a self-assembly method to prepare Pt-ZnIn2S4/RGO/Co3O4-BiVO4 (110) Z-scheme photocatalysts for photocatalytic overall water splitting without sacrificial agents. The decagonal BiVO4 can accumulate holes and electrons on the (110) and (040) sides and top surfaces, respectively, allowing effective spatial separation of photogenerated carriers, so the decagonal BiVO4 with the (110) surface can be used as an oxygen production photocatalyst. ZnIn2S4 was used as a hydrogen production photocatalyst and RGO as an electron mediator to construct Z-scheme heterojunctions with BiVO4 (110). The Z-scheme heterojunction effectively increases the rate of photogenerated charge separation and migration, and the appropriate amount of RGO electron mediators and Pt/Co3O4 co-catalyst loading promotes electron transfer. The material achieves a hydrogen yield of 24.5 μmol g−1 h−1 and an oxygen yield of 11.9 μmol g−1 h−1 under visible-light irradiation in pure water.

Direct Z-Scheme Heterojunction

In 2019, Ding et al. synthesized PtS-ZnIn2S4/WO3-MnO2 direct Z-scheme heterojunctions, achieving a hydrogen yield of 38.8 μmol g−1 h−1 and an oxygen yield of 15.7 μmol g−1 h−1 [91]. They used a hydrothermal method to synthesize orthorhombic WO3•H2O nanoplates, which were dehydrated by solvent heat treatment to transform the orthorhombic phase WO3 to the hexagonal phase. The self-assembly of hexagonal ZnIn2S4 nanosheets on the surface of hexagonal WO3 nanorods to prepare ZnIn2S4/WO3, Pt, and MnO2 were selectively deposited on ZnIn2S4 and WO3 to prepare materials for photocatalytic overall water splitting under pure water without sacrificial agents. Self-assembled ZnIn2S4 on the surface of WO3 nanorods forms an ohmic contact at the interface and a Z-scheme heterojunction between two semiconductor-matched energy band structures (Figure 4d). The construction of Z-scheme heterojunctions effectively promotes the effective spatial separation of photogenerated charges and maintains the high redox capacity of the original semiconductor, resulting in materials with excellent photocatalytic activity.
In 2020, Zhao et al. prepared BiVO4@ZnIn2S4/Ti3C2 MXene quantum dots (BV@ZIS/TC QDs) with layered core–shell structures by in situ growth combined with a two-step solvothermal strategy [93]. They first synthesized BiVO4 microspheres by the hydrothermal method, on which ZnIn2S4 nanosheets were grown in situ to form a hierarchical core–shell structure (Figure 5a). A solid contact surface and matching energy band structure was ensured between two semiconductors to construct the Z-scheme heterojunction, and then it was loaded with Ti3C2 MXene QDs (TC QDs) as a co-catalyst to synthesize composite material for photocatalytic overall water splitting. Material electron transfer pathways are characterized, BiVO4@ZnIn2S4 (BV@ZIS) constructs space-charge-separated all-solid z-structures, and co-catalyst TC QDs form Schottky barriers between the interfaces, effectively promoting photogenerated carrier separation and migration rates (Figure 5b). The unique layered core–shell structure of the material increases light utilization, shortens the charge diffusion distance, increases more active sites, and effectively improves the photocatalytic activity. The material achieves a hydrogen yield of 102.67 μmol g−1 h−1 and an oxygen yield of 50.83 μmol g−1 h−1 under visible-light irradiation.
In 2021, Wang et al. reported a Z-scheme heterojunction ZIS-WO/C-wood (Sv-ZnIn2S4-Ov-WO3/C-wood) catalyst photothermal integrated system for photocatalytic overall water splitting [94]. They first synthesized sulfur-deficient Pt/Sv-ZnIn2S4 (Pt/ZIS) by hydrothermal and reduction methods, followed by oxygen-deficient WO3 (WO) by solvothermal and calcination methods, and CoOx/WO by loading CoOx. Finally, Pt/ZIS and CoOx/WO were dispersed in water to form a solution and spin-coated onto C-wood, and vacuum-assisted heat treatment was applied to construct a photocatalyst photothermal integration system for photocatalytic overall water splitting. ZIS and WO construct Z-scheme heterojunctions, forming a built-in electric field that effectively improves the efficiency of photogenerated charge separation and migration. The conductive material C-wood can act as an electron transfer medium, facilitating electron transfer, and can also use the photothermal effect to convert liquid water to water vapor, transforming the three-phase into a solid/gas two-phase and lowering the carrier recombination and photocatalytic reaction barrier. The photothermal-assisted Z-scheme heterojunction materials effectively increase the photocatalytic activity, achieving a hydrogen yield of 169.2 μmol g−1 h−1 and an oxygen yield of 82.5 μmol g−1 h−1.
Zuo et al. synthesized TiO2-ZnIn2S4 nanoflowers (TNZIS) by integrating ultrathin TiO2 nanosheets into ZnIn2S4 growth solutions in 2021 to construct direct Z-scheme heterojunctions for photocatalytic overall water splitting under co-catalyst-free conditions (Figure 6a) [95]. TiO2 nanosheets (TiO2 NSs) construct direct Z-scheme heterojunctions with ZnIn2S4 to enhance the rate of photogenerated charge separation and inhibit recombination. TiO2 has a larger work function and smaller Fermi energy level than ZnIn2S4 (Figure 6c), causing electrons from ZnIn2S4 to tend to flow into TiO2 to balance the Fermi energy level and bend the energy band, forming a (−)TiO2/(+)ZnIn2S4 built-in electric field at the heterojunction interface (Figure 6b,d) and promoting direct Z-scheme heterojunction formation. The introduction of TiO2 NSs can effectively inhibit the aggregation of ZnIn2S4 materials and increase the specific surface area and active sites of the materials, thus effectively improving the photocatalytic activity. The best material, TNZIS-50, produces 214.9 μmol g−1 h−1 of hydrogen and 81.7 μmol g−1 h−1 of oxygen.
In 2022, Zhang et al. designed and prepared a direct Z-scheme heterojunction BiFeO3/ZnIn2S4 with ferroelectric polarization and internal electric field synergy by a two-step solvothermal method to achieve photocatalytic overall water splitting without a co-catalyst [96]. They prepared BiFeO3 by the solvothermal method and then grew ultrathin ZnIn2S4 nanosheets on BiFeO3 polyhedral particles to prepare BiFeO3/ZnIn2S4. Chalcogenide ABO3-type ferroelectric semiconductors BiFeO3 and ZnIn2S4 construct Z-scheme heterojunctions. Due to the different Fermi energy levels of BiFeO3 and ZnIn2S4 forming a built-in electric field, the ferroelectric polarization and internal electric field effectively promote photogenerated carrier migration and separation efficiency. The increased specific surface area and rich pore distribution of the material provide more catalytic active sites, thus effectively increasing the photocatalyst activity and achieving a hydrogen yield of 87.3 μmol g−1 h−1 and an oxygen yield of 42.3 μmol g−1 h−1.
Yang et al. reported a direct Z-scheme heterojunction BiOBr/ZnIn2S4 (BOB/ZIS) with atomic contact surfaces [97]. They constructed direct Z-scheme heterojunctions with atomic contact surfaces by in situ growth of ZnIn2S4 nanosheets on the surface of BiOBr nanosheets using a solvothermal method to achieve photocatalytic overall water splitting without sacrificial agents and co-catalysts (Figure 7a). As shown in Figure 7b, the disappearance of the Bi-O peak in X-ray photoelectron spectroscopy (XPS) O 1s and the appearance of Bi-S stretching vibrations in the composite at 1116 cm−1 in the FTIR spectrum of BiOBr prove (Figure 7c) that the introduction of TAA into the material during the growth of ZnIn2S4 breaks the Bi-O bond and forms a Bi-S bond at the heterojunction interface, creating an interface with atomic-level seamlessness. Direct Z-scheme heterojunctions with atomic interfacial connections are effective in increasing the photogenerated charge transfer rate, achieving hydrogen yields of 628 μmol g−1 h−1 and oxygen yields of 304 μmol g−1 h−1 under visible-light irradiation in pure water.
In 2022, Zuo et al. synthesized an organic–inorganic hybrid material HC-PDI@ZnIn2S4 OIHs (HPZ) for photocatalytic overall water splitting without co-catalysts and sacrificial agents, achieving hydrogen yields of 275.4 μmol g−1 h−1 and oxygen yields of 138.4 μmol g−1 h−1 under visible-light irradiation [98]. They first synthesized the high-crystalline perylene-dicarboximide supramolecule (HC-PDI) with a highly ordered crystal structure and efficient water oxidation activity by nucleophilic addition and self-assembly, and then synthesized organic–inorganic hybrids by lateral epitaxial growth of ZnIn2S4 nanosheets on highly crystalline HC-PDI nanorods (Figure 7d). As ZnIn2S4 and HC-PDI have different work functions and Fermi energy levels (Figure 7e), electrons in ZnIn2S4 will spontaneously transfer to HC-PDI, creating an electron-consuming layer at the internal interface to generate energy band bending and built-in electric fields, building direct Z-scheme heterojunctions. As shown in Figure 7f, the differential charge density between HC-PDI and ZnIn2S4 can reveal the charge transfer path. The strong covalent coupling between ZnIn2S4 and HC-PDI provides a fast channel for the charge, and the Z-scheme heterojunction effectively promotes the photogenerated charge separation rate. The proper lateral epitaxial growth of ZnIn2S4 gives the composite a layered dendritic structure, which facilitates the improvement of the specific surface area, pore size, pore volume, and visible light collection capacity of the material.
In 2023, Zou et al. constructed Z-scheme heterojunctions with ZnIn2S4 based on the excellent oxidation activity of InVO4 metal oxides [99]. They first synthesized InVO4 nanosheets by the hydrothermal method under acidic conditions, and then formed layered InVO4@ZnIn2S4 (InVZ) heterojunctions by the in situ growth of ZnIn2S4 on them by the magnetic stirring reflux oil bath method. The InVO4 work function is greater than that of ZnIn2S4, causing electrons in ZnIn2S4 to spontaneously enter InVO4, resulting in the bending of the interfacial energy band within the heterojunction and the formation of (−)InVO4@(+)ZnIn2S4. The built-in electric field promotes the construction of Z-scheme heterojunctions. The material semiconductor types and the corresponding conduction and valence band positions were demonstrated by characterization of the Mott–Schottky and Tauc energy band gaps of InVO4 and ZnIn2S4. The construction of InVO4@ZnIn2S4 Z-scheme heterojunctions is further confirmed by their interleaved energy band structures. The Z-heterojunction of InVZ effectively promotes the separation and migration of photogenerated carriers and maintains the high redox activity of the catalyst. The composite material has a larger surface area and pore volume, providing more active sites, thus effectively improving the overall photocatalytic overall water splitting activity of the material. The optimum material InVZ-90 achieves a hydrogen yield of 153.3 μmol g−1 h−1 and an oxygen yield of 76.9 μmol g−1 h−1 under visible light without any co-catalyst.

4.2.2. Schottky Junctions

The Schottky junction is an interface between a metal and a semiconductor. As the metal and semiconductor have different escape work and Fermi energy levels, the Fermi energy level shifts when the metal and semiconductor come into contact until the Fermi energy level equilibrates [100,101]. When the semiconductor is an n-type semiconductor and the metal escape work is greater than the semiconductor Fermi energy level, a Schottky barrier will form at the interface, limiting the flow of electrons from the semiconductor to the metal, inhibiting photogenerated carrier compounding, and effectively improving photocatalytic activity.
In 2021, Cai et al. synthesized a yolk–shell ZnIn2S4-based photocatalyst NiCo2S4/ZnIn2S4/Co3O4 for photocatalytic overall water splitting without sacrificial agents [102]. They first prepared nickel–cobalt-based metal–organic framework (MOF) materials as precursors by the solvothermal method, synthesized yolk–shell NiCo2S4 (NCS) with semi-metallic properties by sulfidation reaction and heat treatment, then grew ZnIn2S4 nanosheets in situ on the surface of NCS by low-temperature solvothermal method, and finally decorated Co3O4 nanoparticles to prepare photocatalysts (Figure 8a). The characteristic peak of NiCo-glyceric acid disappeared and shifted to NiCo2S4 when vulcanized at different vulcanization temperatures for 8 h. The precursors were vulcanized at 150, 180, and 210 °C for 8 h to form ball-in-ball hollow spheres, yolk–shell hollow spheres, and a single hollow sphere, respectively, with the yolk–shell structure providing more active sites and improving the photocatalytic activity of the material. As shown in Figure 8b,c, based on the energy band diagrams of ZnIn2S4 and NCS as well as the energy band diagrams after contact, it is shown that the semi-metallic NCS forms a Schottky-specific heterojunction with ZnIn2S4 to facilitate charge transfer. The photocatalyst has a unique yolk–shell structure that locates reduction and oxidation sites on the inner and outer surfaces of ZnIn2S4, respectively, allowing for directional charge separation, inhibition of inverse reactions, and providing more active sites.
In 2022, Liu et al. designed and prepared a sandwich structure material Nb4C3Tx MXene@ZnIn2S4-OH by in situ growth and peroxide plasma post-treatment [103]. They first prepared uniformly dispersed accordion-like multilayer Nb4C3Tx MXenes by selective chemical etching, then ultrathin ZnIn2S4 was epitaxially grown on their surface to synthesize the sandwich structure Nb4C3Tx MXene@ ZnIn2S4 at the double heterojunction interface (Figure 9a), followed by a peroxy plasma technique to generate many hydroxyl functional groups to obtain photocatalysts for photocatalytic monolithic water splitting in the absence of sacrificial agents. The work function of ZnIn2S4 and Nb4C3Tx MXene shows that the transfer of electrons from ZnIn2S4 to Nb4C3Tx MXene at the interface causes the ZnIn2S4 energy band to rise, holes remain in ZnIn2S4, and the Nb4C3Tx MXene@ZnIn2S4 photocatalyst forms a Schottky junction (Figure 9b,c). The photocatalyst promotes photogenerated carrier transfer and has a larger surface area and pore size and pore volume for rapid adsorption of water molecules. In addition, the OH functional group on the surface of ZnIn2S4 collects photogenerated holes, and the unique eggshell-type structure of the composite material allows the generated H2 and O2 to be distributed internally and externally, achieving spatial separation of photogenerated carriers and suppressing the inverse reaction. The photocatalyst achieves a hydrogen yield of 53.8 μmol g−1 h−1 and an oxygen yield of 26.7 μmol g−1 h−1 under visible-light irradiation.

4.3. Loaded Co-Catalyst

Photocatalytic overall water splitting is mainly based on three basic processes of photocatalysis. The low overall photogenerated charge separation and migration efficiency and the severe compounding of photogenerated charges limit the catalytic photocatalytic activity due to the Coulombic force constraint and the high potential barrier of the transfer process, as well as the unfavorable thermodynamic process of water generation of H2 and O2. In this regard, the loading of co-catalysis on semiconductors is a preferable strategy. When a semiconductor is loaded with a co-catalyst, photogenerated electrons and photogenerated holes migrate to the co-catalyst for reduction–oxidation reactions, effectively facilitating photogenerated charge separation and migration. In addition, loaded co-catalysts have the advantage of reducing the activation energy, inhibiting photo-corrosion, and providing an abundance of surface reaction sites [104,105]. Unlike modification strategies such as doping and the construction of heterojunctions, catalysts loaded with co-catalysts can usually be synthesized using photo-deposition.
Suitable co-catalysts are available for both the hydrogen and oxygen production halves of the photocatalytic overall water splitting. The main hydrogen-dissolving co-catalysts are noble metals such as Pt and Au, transition metal monomers such as Co and Ni, transition metal sulfides such as NiP and MoS2, and phosphides. The main oxygen-dissolving co-catalysts are noble metal oxides such as RuO2 and IrO2, and transition metal oxides such as CoOx [106]. In photocatalytic overall water splitting systems, we usually load hydrogen- and oxygen-precipitating co-catalysts onto the catalyst separately to reduce the catalyst surface potential barrier, improve the efficiency of photogenerated charge separation and migration, and better achieve photocatalytic overall water splitting. For example, CdS loaded with Pt and Ru complexes as hydrogen and oxygen production co-catalysts, Mg-doped BaTaO2N loaded with Cr2O3/(Na)Rh and IrO2 as hydrogen and oxygen production co-catalysts, and Ge3N4 loaded with spatially separated mixed cathode and anode co-catalysts, and CoOx-Mo2N all enable photocatalytic monolithic water splitting [107,108].
The photocatalytic activity can be influenced by adjusting the amount of co-catalyst loading, the type of co-catalyst, the shape of the co-catalyst, etc. In 2023, Jing et al. prepared spatially separated double-coordinated co-catalyst ZnIn2S4 composites [109]. They first prepared the hexagonal crystalline phase ZnIn2S4 by the hydrothermal method, and then prepared ZnIn2S4-Pt-Cr and ZnIn2S4-Rh-Cr by photo-deposition with Pt-Cr and Rh-Cr co-catalysts loaded on the surface of ZnIn2S4. The elements Pt and Rh are used as water reduction sites and Cr as water oxidation sites to promote hydrogen and oxygen production reactions, respectively, for overall water splitting. The energy-dispersive X-ray spectroscopy (EDS)-scanned elemental signals of Rh, Cr, and Pt elements at different positions to probe the loading position of the co-catalysts showed that: Pt and Cr elements had consistent signal changes and were hybrids; Rh and Cr elements had different signal changes and achieved spatial separation. In addition, X-ray photoelectron spectroscopy (XPS) demonstrated the formation of Rh-S bonds in the ZnIn2S4-Rh-Cr material, facilitating the spatial separation of the co-catalyst. The materials loaded with double-assisted catalysts can effectively improve the separation and migration efficiency of photogenerated carriers, inhibit photo-corrosion of the materials, and improve the photocatalytic performance of the materials. The photocatalytic overall water splitting performance of the loaded spatially separated dual-assisted catalyst ZnIn2S4-Rh-Cr composite was twice that of the ZnIn2S4-Pt-Cr material, achieving hydrogen yields of 5.9 μmol h−1 and oxygen yields of 2.9 μmol h−1 at AM 1.5 G. As shown in Table 2, in recent years, researchers have achieved overall water splitting of ZnIn2S4-based photocatalysts through modification strategies such as doping, vacancies, construction of heterojunctions, and loading of co-catalysts.

5. Conclusions and Outlook

In a two-carbon context, the use of clean, renewable solar energy is an ideal solution to current energy and environmental problems. Photocatalytic overall water splitting, which produces clean, high-energy-density hydrogen without sacrificial agents, is the ideal, low-cost method for solar–chemical energy conversion in a variety of solar applications. Compared with oxides and nitrides, ZnIn2S4 as a ternary metal sulfide has the advantages of tunable band gap, satisfying the energy band of photocatalytic overall water splitting, strong photostability, and easy preparation, which makes it a more ideal material for photocatalytic overall water splitting. In this study, we systematically review recent advances in ZnIn2S4-based photocatalysts for photocatalytic overall water splitting in solar–chemical energy conversion. This study mainly introduces the basic principles of its photocatalytic overall water splitting, the properties of the ZnIn2S4 photocatalyst, including its crystal configuration and energy band structure, the main synthesis methods, and morphology. It also reviews the research progress of photocatalytic overall water splitting of ZnIn2S4-based photocatalytic materials, including modification strategies such as elemental doping, vacancy defects, the construction of heterojunctions, and the loading of co-catalysts.
Although good progress has been made in the current study of ZnIn2S4-based photocatalysts for water splitting, there are still some limitations of ZnIn2S4-based photocatalysts for water splitting that limit their wide application. The overall photogenerated charge separation and utilization efficiency are low due to the Coulombic force constraint between the photogenerated electrons and holes and the interfacial potential barriers between charge transfers, which are much higher than the photogenerated electrons and holes that they separate and migrate to the redox sites to participate in the reaction. Furthermore, although ZnIn2S4 has a good light collection capacity, it is mainly responsive to visible light and does not utilize the near-infrared light, which is nearly 50% of solar energy, and the inevitable existence of ZnIn2S4-based photocatalysts as metal sulfides with their photo-corrosive fluxes affects their photocatalytic overall water splitting activity. Therefore, more efforts are still needed to make ZnIn2S4-based photocatalysts better for use in photocatalytic overall water splitting systems:
(1) Building active crystal surfaces. Very few studies have systematically addressed the relationship between the active crystal plane of ZnIn2S4 photocatalysts and the overall water splitting activity of photocatalysis. Usually, photocatalytic reactions occur on the surface of the material, so the specific surface area of the material and the exposed crystalline surface all affect the catalyst activity. Particularly due to the anisotropy of crystals, different crystal faces have different structures and atomic arrangements and usually exhibit different properties. It is important to clarify the relationship between different crystal types and their different crystal faces and photocatalytic activity, and to explore which crystal exposures alter the conduction/valence band position of the material and enhance electron reduction/hole oxidation, and which crystal exposures accumulate photogenerated charges/holes. Improving the photogenerated charge separation and migration efficiency and reducing photo-corrosion can be achieved by modulating the crystalline surface of ZnIn2S4-based photocatalysts. Therefore, if the relationship between the different crystalline facets of ZnIn2S4 and photocatalytic activity can be clarified, we can better design photocatalytic materials with higher activity and significantly improve their photocatalytic overall water splitting performance.
(2) Photothermal assistance for broad spectrum utilization. Most current applications for photocatalytic monolithic water splitting ZnIn2S4-based materials respond to visible light and largely fail to utilize near-infrared light, which accounts for approximately 50% of sunlight; therefore, expanding the efficiency of light utilization is essential to achieve high-efficiency photocatalytic overall water splitting. Due to the inherent nature of ZnIn2S4 materials, most current ZnIn2S4-based photocatalytic overall water splitting systems employ very limited modification strategies such as doping, vacancies, and the construction of heterojunctions to extend the photo-response range. The combination of photothermal assistance with ZnIn2S4-based photocatalysts is a very promising strategy for enhancing overall water splitting due to the advantages of photothermal catalysis in terms of wider spectral utilization, enhanced free carrier concentration, and promotion of photogenerated carrier separation. Further efforts are still needed to explore the design of photothermal-assisted ZnIn2S4-based photocatalysts to achieve broad-spectrum photocatalytic total hydrolysis and the development of low-sink cost materials to replace expensive precious metals.
(3) Morphological adjustments. By adjusting the morphology of ZnIn2S4-based photocatalytic materials, we can increase the specific surface area, pore size, and volume, and increase the active site to improve the light collection capacity, such as hollow structures for broadband absorption and ultra-thin nanosheets to shorten the photogenerated charge migration efficiency. The current sandwich structure and yolk–shell structure in ZnIn2S4-based photocatalytic overall water splitting applications have spatially separated water redox sites, allowing for directional separation of photogenerated charges, improved photocatalytic activity, as well as a layered core–shell structure to enhance incident light scattering and reflection effects and improve light utilization efficiency. Although the current modified morphologies are all very good at improving the overall photocatalytic overall water splitting activity, there is still room for improvement. Further research is needed to determine whether some of the complex morphological synthesis steps and materials used can be simplified and cost-effective, whether the prepared materials are stable, and whether morphological adjustments can be made to the ZnIn2S4 native catalyst to achieve single-component photocatalytic overall water splitting.

Author Contributions

Writing—original draft preparation, Y.Y.; visualization, Z.C.; project administration, X.C.; writing—review and editing, W.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China: 22006057 and 51902140. Thanks to the Instrumental Analysis Center, Jiangsu University of Science and Technology.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tao, X.; Zhao, Y.; Wang, S.; Li, C.; Li, R. Recent advances and perspectives for solar-driven water splitting using particulate photocatalysts. Chem. Soc. Rev. 2022, 51, 3561–3608. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, Q.; Domen, K. Particulate Photocatalysts for Light-Driven Water Splitting: Mechanisms, Challenges, and Design Strategies. Chem. Rev. 2020, 120, 919–985. [Google Scholar] [CrossRef] [PubMed]
  3. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  4. Hu, J.; Chen, C.; Zheng, Y.; Zhang, G.; Li, C.M. Spatially Separating Redox Centers on Zlog cheme ZnIn2S4/BiVO4 Hierarchical Heterostructure for Highly Efficient Photocatalytic Hydrogen Evolution. Small 2020, 16, 2002988. [Google Scholar] [CrossRef]
  5. Su, T.; Men, C.; Chen, L.; Chu, B.; Luo, X.; Ji, H.; Chen, J.; Qin, Z. Sulfur Vacancy and Ti3C2Tx Cocatalyst Synergistically Boosting Interfacial Charge Transfer in 2D/2D Ti3C2Tx/ZnIn2S4 Heterostructure for Enhanced Photocatalytic Hydrogen Evolution. Adv. Sci. 2022, 9, 2103715. [Google Scholar] [CrossRef]
  6. Du, C.; Zhang, Q.; Lin, Z.; Yan, B.; Xia, C.; Yang, G. Half-unit-cell ZnIn2S4 monolayer with sulfur vacancies for photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2019, 248, 193–201. [Google Scholar] [CrossRef]
  7. Liu, K.; Zhang, B.; Zhang, J.; Lin, W.; Wang, J.; Xu, Y.; Xiang, Y.; Hisatomi, T.; Domen, K.; Ma, G. Synthesis of Narrow-Band-Gap GaN:ZnO Solid Solution for Photocatalytic Overall Water Splitting. ACS Catal. 2022, 12, 14637–14646. [Google Scholar] [CrossRef]
  8. Zhang, G.G.; Wang, X.C. Oxysulfide Semiconductors for Photocatalytic Overall Water Splitting with Visible Light. Angew. Chem.-Int. Ed. 2019, 58, 15580–15582. [Google Scholar] [CrossRef]
  9. Gogoi, D.; Shah, A.; Rambabu, P.; Qureshi, M.; Golder, A.; Peela, N. Step-Scheme Heterojunction between CdS Nanowires and Facet-Selective Assembly of MnOx-BiVO4 for an Efficient Visible-Light-Driven Overall Water Splitting. ACS Appl. Mater. Interfaces 2021, 13, 45475–45487. [Google Scholar] [CrossRef]
  10. Iwashina, K.; Iwase, A.; Ng, Y.H.; Amal, R.; Kudo, A. Z-schematic water splitting into H2 and O2 using metal sulfide as a hydrogen-evolving photocatalyst and reduced graphene oxide as a solid-state electron mediator. J. Am. Chem. Soc. 2015, 137, 604–607. [Google Scholar] [CrossRef]
  11. Hayat, A.; Sohail, M.; Anwar, U.; Taha, T.A.; El-Nasser, K.S.; Alenad, A.M.; Al-Sehemi, A.G.; Alghamdi, N.A.; Al-Hartomy, O.A.; Amin, M.A.; et al. Enhanced photocatalytic overall water splitting from an assembly of donor-π-acceptor conjugated polymeric carbon nitride. J. Colloid Interface Sci. 2022, 624, 411–422. [Google Scholar] [CrossRef]
  12. Sun, S.; Gao, R.; Liu, X.; Pan, L.; Shi, C.; Jiang, Z.; Zhang, X.; Zou, J.J. Engineering interfacial band bending over bismuth vanadate/carbon nitride by work function regulation for efficient solar-driven water splitting. Sci. Bull. 2021, 67, 389–397. [Google Scholar] [CrossRef]
  13. Pan, Y.; Yuan, X.; Jiang, L.; Yu, H.; Zhang, J.; Wang, H.; Guan, R.; Zeng, G. Recent advances in synthesis, modification and photocatalytic applications of micro/nano-structured zinc indium sulfide. Chem. Eng. J. 2018, 354, 407–431. [Google Scholar] [CrossRef]
  14. Zhu, Q.; Xu, Q.; Du, M.; Zeng, X.; Zhong, G.; Qiu, B.; Zhang, J. Recent Progress of Metal Sulfide Photocatalysts for Solar Energy Conversion. Adv. Mater. 2022, 34, e2202929. [Google Scholar] [CrossRef]
  15. Hao, X.Q.; Zhou, J.; Cui, Z.W.; Wang, Y.C.; Wang, Y.; Zou, Z. Zn-vacancy mediated electron-hole separation in ZnS/g-C3N4 heterojunction for efficient visible-light photocatalytic hydrogen production. Appl. Catal. B Environ. 2018, 229, 41–51. [Google Scholar] [CrossRef]
  16. Xue, S.; Huang, W.; Lin, W.; Xing, W.; Shen, M.; Ye, X.; Liang, X.; Yang, C.; Hou, Y.; Yu, Z. Interfacial engineering of lattice coherency at ZnO-ZnS photocatalytic heterojunctions—ScienceDirect. Chem Catal. 2022, 2, 125–139. [Google Scholar] [CrossRef]
  17. Yu, J.; Yu, Y.; Peng, Z.; Wei, X.; Bei, C. Morphology-dependent photocatalytic H2-production activity of CdS. Appl. Catal. B Environ. 2014, 156–157, 184–191. [Google Scholar] [CrossRef]
  18. Yuan, Y.-J.; Chen, D.; Yu, Z.-T.; Zou, Z.-G. Cadmium sulfide-based nanomaterials for photocatalytic hydrogen production. J. Mater. Chem. A Mater. Energy Sustain. 2018, 6, 11606–11630. [Google Scholar] [CrossRef]
  19. Ren, Y.; Foo, J.J.; Zeng, D.; Ong, W.J. ZnIn2S4-Based Nanostructures in Artificial Photosynthesis: Insights into Photocatalytic Reduction toward Sustainable Energy Production. Small Struct. 2022, 3, 2200017. [Google Scholar] [CrossRef]
  20. Chen, K.; Shi, Y.; Shu, P.; Luo, Z.; Shi, W.; Guo, F. Construction of core–shell FeS2@ZnIn2S4 hollow hierarchical structure S-scheme heterojunction for boosted photothermal-assisted photocatalytic H2 production. Chem. Eng. J. 2023, 454, 140053. [Google Scholar] [CrossRef]
  21. Zhang, G.; Wu, H.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. A mini-review on ZnIn2S4-Based photocatalysts for energy and environmental application. Green Energy Environ. 2022, 7, 176–204. [Google Scholar] [CrossRef]
  22. Chandrasekaran, S.; Yao, L.; Deng, L.; Bowen, C.; Zhang, Y.; Chen, S.; Lin, Z.; Peng, F.; Zhang, P. Recent advances in metal sulfides: From controlled fabrication to electrocatalytic, photocatalytic and photoelectrochemical water splitting and beyond. Chem. Soc. Rev. 2019, 48, 4178–4280. [Google Scholar] [CrossRef] [PubMed]
  23. Cai, Y.; Shi, Y.; Shi, W.; Bai, S.; Yang, S.; Guo, F. A one-photon excitation pathway in 0D/3D CoS2/ZnIn2S4 composite with nanoparticles on micro-flowers structure for boosted visible-light-driven photocatalytic hydrogen evolution. Compos. Part B Eng. 2022, 238, 109955. [Google Scholar] [CrossRef]
  24. Shi, W.; Hao, C.; Fu, Y.; Guo, F.; Tang, Y.; Yan, X. Enhancement of synergistic effect photocatalytic/persulfate activation for degradation of antibiotics by the combination of photo-induced electrons and carbon dots. Chem. Eng. J. 2022, 433, 133741. [Google Scholar] [CrossRef]
  25. Liu, C.; Zhang, Q.; Zou, Z. Recent advances in designing ZnIn2S4-based heterostructured photocatalysts for hydrogen evolution. J. Mater. Sci. Technol. 2023, 139, 167–188. [Google Scholar] [CrossRef]
  26. Yadav, G.; Ahmaruzzaman, M. Recent progress on synthesis and modifications of ZnIn2S4 based novel hybrid materials for potential applications. Mater. Sci. Eng. B 2023, 292, 116418. [Google Scholar] [CrossRef]
  27. Wang, Z.; Li, C.; Domen, K. Recent developments in heterogeneous photocatalysts for solar-driven overall water splitting. Chem. Soc. Rev. 2019, 48, 2109–2125. [Google Scholar] [CrossRef]
  28. He, R.; Ran, J. Dilemma faced by photocatalytic overall water splitting. J. Mater. Sci. Technol. 2023, 157, 107–109. [Google Scholar] [CrossRef]
  29. Navalon, S.; Dhakshinamoorthy, A.; Alvaro, M.; Ferrer, B.; Garcia, H. Metal-Organic Frameworks as Photocatalysts for Solar-Driven Overall Water Splitting. Chem. Rev. 2023, 123, 445–490. [Google Scholar] [CrossRef]
  30. Bie, C.; Wang, L.; Yu, J. Challenges for photocatalytic overall water splitting. Chem 2022, 8, 1567–1574. [Google Scholar] [CrossRef]
  31. Huang, Y.; Li, D.; Feng, S.; Jia, Y.; Guo, S.; Wu, X.; Chen, M.; Shi, W. Pt Atoms/Clusters on Ni-phytate-sensitized Carbon Nitride for Enhanced NIR-light-driven Overall Water Splitting beyond 800 nm. Angew. Chem. Int. Ed. Engl. 2022, 61, e202212234. [Google Scholar] [CrossRef]
  32. Zhou, P.; Navid, I.A.; Ma, Y.; Xiao, Y.; Wang, P.; Ye, Z.; Zhou, B.; Sun, K.; Mi, Z. Solar-to-hydrogen efficiency of more than 9% in photocatalytic water splitting. Nature 2023, 613, 66–70. [Google Scholar] [CrossRef]
  33. Hu, L.; Huang, J.; Wang, J.; Jiang, S.; Sun, C.; Song, S. Efficiently photocatalytic H2O overall splitting within the strengthened polarized field by reassembling surface single atoms. Appl. Catal. B Environ. 2023, 320, 121945. [Google Scholar] [CrossRef]
  34. Chen, J.; Xin, F.; Yin, X.; Xiang, T.; Wang, Y. Synthesis of hexagonal and cubic ZnIn2S4 nanosheets for the photocatalytic reduction of CO2 with methanol. RSC Adv. 2015, 5, 3833–3839. [Google Scholar] [CrossRef]
  35. Lee, J.; Kim, H.; Lee, T.; Jang, W.; Lee, K.H.; Soon, A. Revisiting Polytypism in Hexagonal Ternary Sulfide ZnIn2S4 for Photocatalytic Hydrogen Production within the Z-Scheme. Chem. Mater. 2019, 31, 9148–9155. [Google Scholar] [CrossRef]
  36. Yue, Y.; Zou, J. One-step hydrothermal synthesis of rhombohedral ZnIn2S4 with high visible photocatalytic activity for aqueous pollutants removal. J. Colloid Interface Sci. 2023, 640, 270–280. [Google Scholar] [CrossRef]
  37. Shen, S.; Guo, P.; Zhao, L.; Du, Y.; Guo, L. Insights into photoluminescence property and photocatalytic activity of cubic and rhombohedral ZnIn2S4. J. Solid State Chem. 2011, 184, 2250–2256. [Google Scholar] [CrossRef]
  38. Wang, J.; Sun, S.; Zhou, R.; Li, Y.; He, Z.; Ding, H.; Chen, D.; Ao, W. A review: Synthesis, modification and photocatalytic applications of ZnIn2S4. J. Mater. Sci. Technol. 2021, 78, 1–19. [Google Scholar] [CrossRef]
  39. Guo, X.; Liu, X.; Wang, M.; Yan, J.; Chen, Y.; Liu, S. Unveiling the Origin of Co3O4 Quantum Dots for Photocatalytic Overall Water Splitting. Small 2023, 19, e2206695. [Google Scholar] [CrossRef]
  40. Peng, S.; Li, L.; Wu, Y.; Jia, L.; Tian, L.; Srinivasan, M.; Ramakrishna, S.; Yan, Q.; Mhaisalkar, S.G. Size- and shape-controlled synthesis of ZnIn2S4 nanocrystals with high photocatalytic performance. CrystEngComm 2013, 15, 1922–1930. [Google Scholar] [CrossRef]
  41. Shi, L.; Yin, P.; Dai, Y. Synthesis and photocatalytic performance of ZnIn2S4 nanotubes and nanowires. Langmuir 2013, 29, 12818–12822. [Google Scholar] [CrossRef] [PubMed]
  42. Ambrosi, A.; Pumera, M. Templated Electrochemical Fabrication of Hollow Molybdenum Sulfide Microstructures and Nanostructures with Catalytic Properties for Hydrogen Production. ACS Catal. 2016, 6, 3985–3993. [Google Scholar] [CrossRef]
  43. Luo, R.-B.; Liu, Q.-J.; Fan, D.-H.; Liu, Z.-T. First-principles calculations to investigate structural, electronic and optical properties of two-dimensional ZnIn2S4. Appl. Surf. Sci. 2022, 605, 154739. [Google Scholar] [CrossRef]
  44. Zhang, S.; Liu, X.; Liu, C.; Luo, S.; Wang, L.; Cai, T.; Zeng, Y.; Yuan, J.; Dong, W.; Pei, Y.; et al. MoS2 Quantum Dot Growth Induced by S Vacancies in a ZnIn2S4 Monolayer: Atomic-Level Heterostructure for Photocatalytic Hydrogen Production. ACS Nano 2018, 12, 751–758. [Google Scholar] [CrossRef] [PubMed]
  45. Shen, J.; Zai, J.; Yuan, Y.; Qian, X. 3D hierarchical ZnIn2S4: The preparation and photocatalytic properties on water splitting. Int. J. Hydrogen Energy 2012, 37, 16986–16993. [Google Scholar] [CrossRef]
  46. Hu, X.; Yu, J.C.; Gong, J.; Li, Q. Rapid Mass Production of Hierarchically Porous ZnIn2S4 Submicrospheres via a Microwave-Solvothermal Process. Cryst. Growth Des. 2007, 7, 2444–2448. [Google Scholar] [CrossRef]
  47. Fang, F.; Chen, L.; Chen, Y.B.; Wu, L.M. Synthesis and Photocatalysis of ZnIn2S4 Nano/Micropeony. J. Phys. Chem. C 2010, 114, 2393–2397. [Google Scholar] [CrossRef]
  48. Zhang, E.; Zhu, Q.; Huang, J.; Liu, J.; Tan, G.; Sun, C.; Li, T.; Liu, S.; Li, Y.; Wang, H.; et al. Visually resolving the direct Z-scheme heterojunction in CdS@ZnIn2S4 hollow cubes for photocatalytic evolution of H2 and H2O2 from pure water. Appl. Catal. B Environ. 2021, 293, 120213. [Google Scholar] [CrossRef]
  49. Chaudhari, N.S.; Warule, S.S.; Kale, B.B. Architecture of rose and hollow marigold-like ZnIn2S4 flowers: Structural, optical and photocatalytic study. RSC Adv. 2014, 4, 12182–12187. [Google Scholar] [CrossRef]
  50. Chen, Z.X.; Li, D.Z.; Zhang, W.J.; Chen, C.; Li, W.J.; Sun, M.; He, Y.H.; Fu, X.Z. Low-Temperature and Template-Free Synthesis of ZnIn2S4 Microspheres. Inorg. Chem. 2010, 40, 9766–9772. [Google Scholar] [CrossRef]
  51. Ding, S.; Liu, X.; Shi, Y.; Liu, Y.; Zhou, T.; Guo, Z.; Hu, J. Generalized Synthesis of Ternary Sulfide Hollow Structures with Enhanced Photocatalytic Performance for Degradation and Hydrogen Evolution. ACS Appl. Mater. Interfaces 2018, 10, 17911–17922. [Google Scholar] [CrossRef] [PubMed]
  52. Gou, X.; Cheng, F.; Shi, Y.; Zhang, L.; Shen, P. Shape-controlled synthesis of ternary chalcogenide ZnIn2S4 and CuIn(S,Se)2 nano-/microstructures via facile solution route. J. Am. Chem. Soc. 2006, 128, 7222–7229. [Google Scholar] [CrossRef] [PubMed]
  53. Chen, Y.; Huang, R.; Chen, D.; Wang, Y.; Liu, W.; Li, X.; Li, Z. Exploring the different photocatalytic performance for dye degradations over hexagonal ZnIn2S4 microspheres and cubic ZnIn2S4 nanoparticles. ACS Appl. Mater. Interfaces 2012, 4, 2273–2279. [Google Scholar] [CrossRef] [PubMed]
  54. Yin, X.; Lv, B.; Kang, Y.; Xu, X.; Lei, X.; Li, L.; Wang, H.; Xi, H.; Yang, J.; Yang, Z. Solvent-Induced ZnIn2S4 Nanosheets Self-assembled Micro-Flowers to Boosting the Photocatalytic Semi-dehydrogenation of 1,2,3,4-Tetrahydroisoquinoline. Catal. Lett. 2022, 153, 570–583. [Google Scholar] [CrossRef]
  55. Zou, P.; Su, G.; Li, Z.; Li, Y.; Zhou, T.; Kang, Y. Oxalic acid modified hexagonal ZnIn2S4 combined with bismuth oxychloride to fabricate a hierarchical dual Z-scheme heterojunction: Accelerating charge transfer to improve photocatalytic activity. Colloids Surf. A Physicochem. Eng. Asp. 2022, 633, 127835. [Google Scholar] [CrossRef]
  56. Su, L.; Ye, X.; Meng, S.; Fu, X.; Chen, S. Effect of different solvent on the photocatalytic activity of ZnIn2S4 for selective oxidation of aromatic alcohols to aromatic aldehydes under visible light irradiation. Appl. Surf. Sci. 2016, 384, 161–174. [Google Scholar] [CrossRef]
  57. Shen, S.; Zhao, L.; Guo, L. Morphology, structure and photocatalytic performance of ZnIn2S4 synthesized via a solvothermal/hydrothermal route in different solvents. J. Phys. Chem. Solids 2008, 69, 2426–2432. [Google Scholar] [CrossRef]
  58. Salavati-Niasari, M.; Ranjbar, M.; Sabet, M. Synthesis and Characterization of ZnIn2S4 Nanoparticles by a Facile Microwave Approach. J. Inorg. Organomet. Polym. Mater. 2012, 23, 452–457. [Google Scholar] [CrossRef]
  59. Yang, W.; Liu, B.; Fang, T.; Jennifer, W.A.; Christophe, L.; Li, Z.; Zhang, X.; Jiang, X. Layered crystalline ZnIn2S4 nanosheets: CVD synthesis and photo-electrochemical properties. Nanoscale 2016, 8, 18197–18203. [Google Scholar] [CrossRef]
  60. Hemalatha, S.; Illakkiya, J.T.; Oommen, R.; Rajalakshmi, P.U. Marigold microstructure of zinc thioindate (ZnIn2S4) thin film—Characteristics. Optik 2016, 127, 3858–3861. [Google Scholar] [CrossRef]
  61. Huang, J.; Cheuk, W.; Wu, Y.; Lee, F.; Ho, W. Template-free synthesis of ternary sulfides submicrospheres as visible light photocatalysts by ultrasonic spray pyrolysis. Catal. Sci. Technol. 2012, 2, 1825. [Google Scholar] [CrossRef]
  62. Cadiş, A.I.; Mureșan, L.E.; Perhaița, I.; Pop, L.C.; Saszet, K.; Barbu-Tudoran, L.; Borodi, G. Peculiarities on methyl orange adsorption by porous ZnIn2S4 prepared in different conditions. J. Nanoparticle Res. 2022, 24, 1–13. [Google Scholar] [CrossRef]
  63. Zhang, Y.C.; Afzal, N.; Pan, L.; Zhang, X.; Zou, J.J. Structure-Activity Relationship of Defective Metal-Based Photocatalysts for Water Splitting: Experimental and Theoretical Perspectives. Adv. Sci. 2019, 6, 1900053. [Google Scholar] [CrossRef] [Green Version]
  64. Su, Z.; Fang, F.; Li, X.; Han, W.; Liu, X.; Chang, K. Synergistic surface oxygen defect and bulk Ti3+ defect engineering on SrTiO3 for enhancing photocatalytic overall water splitting. J. Colloid Interface Sci. 2022, 626, 662–673. [Google Scholar] [CrossRef]
  65. Jiang, L.; Yuan, X.; Pan, Y.; Liang, J.; Zeng, G.; Wu, Z.; Wang, H. Doping of graphitic carbon nitride for photocatalysis: A reveiw. Appl. Catal. B Environ. 2017, 217, 388–406. [Google Scholar] [CrossRef]
  66. Ma, J.; Long, R.; Liu, D.; Low, J.; Xiong, Y. Defect Engineering in Photocatalytic Methane Conversion. Small Struct. 2021, 3, 2100147. [Google Scholar] [CrossRef]
  67. Patnaik, P. Recent advances in anion doped g-C3N4 photocatalysts: A review. Carbon 2021, 172, 682–711. [Google Scholar] [CrossRef]
  68. Yang, R.; Mei, L.; Fan, Y.; Zhang, Q.; Zhu, R.; Amal, R.; Yin, Z.; Zeng, Z. ZnIn2S4-Based Photocatalysts for Energy and Environmental Applications. Small Methods 2021, 5, e2100887. [Google Scholar] [CrossRef]
  69. Pan, R.; Hu, M.; Liu, J.; Li, D.; Wan, X.; Wang, H.; Li, Y.; Zhang, X.; Wang, X.; Jiang, J.; et al. Two-Dimensional All-in-One Sulfide Monolayers Driving Photocatalytic Overall Water Splitting. Nano Lett. 2021, 21, 6228–6236. [Google Scholar] [CrossRef]
  70. Sun, B.; Bu, J.; Chen, X.; Fan, D.; Li, S.; Li, Z.; Zhou, W.; Du, Y. In-situ interstitial zinc doping-mediated efficient charge separation for ZnIn2S4 nanosheets visible-light photocatalysts towards optimized overall water splitting. Chem. Eng. J. 2022, 435, 135074. [Google Scholar] [CrossRef]
  71. Sun, B.; Fan, D.; Chen, X.; Li, Z.; Zhou, W.; Du, Y. Heteroatom-induced domain electrostatic potential difference in ZnIn2S4 nanosheets for efficient charge separation and boosted photocatalytic overall water splitting. Mater. Chem. Front. 2022, 6, 1795–1802. [Google Scholar] [CrossRef]
  72. Jing, H.; Ren, J.; Yue, J.; Liu, S.; Liang, Q.; Wu, R.; Wang, Y.; Fang, Z.; Li, H.; Wei, S. ZnIn2S4 with oxygen atom doping and surface sulfur vacancies for overall water splitting under visible light irradiation. Catal. Sci. Technol. 2023, 13, 226–232. [Google Scholar] [CrossRef]
  73. Zhang, S.; Yi, X.; Hu, G.; Chen, M.; Shen, H.; Li, B.; Yang, L.; Dai, W.; Zou, J.; Luo, S. Configuration regulation of active sites by accurate doping inducing self-adapting defect for enhanced photocatalytic applications: A review. Coord. Chem. Rev. 2023, 478, 214970. [Google Scholar] [CrossRef]
  74. He, J.; Hu, L.; Shao, C.; Jiang, S.; Sun, C.; Song, S. Photocatalytic H2O Overall Splitting into H2 Bubbles by Single Atomic Sulfur Vacancy CdS with Spin Polarization Electric Field. ACS Nano 2021, 15, 18006–18013. [Google Scholar] [CrossRef]
  75. Jing, H.; Xu, G.; Yao, B.; Ren, J.; Wang, Y.; Fang, Z.; Liang, Q.; Wu, R.; Wei, S. Sulfur Vacancy-Enriched Rhombohedral ZnIn2S4 Nanosheets for Highly Efficient Photocatalytic Overall Water Splitting under Visible Light Irradiation. ACS Appl. Energy Mater. 2022, 5, 10187–10195. [Google Scholar] [CrossRef]
  76. Abdul Nasir, J.; Munir, A.; Ahmad, N.; Haq, T.U.; Khan, Z.; Rehman, Z. Photocatalytic Z-Scheme Overall Water Splitting: Recent Advances in Theory and Experiments. Adv. Mater. 2021, 33, 2105195. [Google Scholar] [CrossRef]
  77. Lu, C.; Guo, F.; Yan, Q.; Zhang, Z.; Li, D.; Wang, L.; Zhou, Y. Hydrothermal synthesis of type II ZnIn2S4/BiPO4 heterojunction photocatalyst with dandelion-like microflower structure for enhanced photocatalytic degradation of tetracycline under simulated solar light. J. Alloys Compd. 2019, 811, 151976. [Google Scholar] [CrossRef]
  78. Shi, Y.; Li, L.; Xu, Z.; Guo, F.; Shi, W. Construction of full solar-spectrum available S-scheme heterojunction for boosted photothermal-assisted photocatalytic H2 production. Chem. Eng. J. 2023, 459, 151976. [Google Scholar] [CrossRef]
  79. Wang, J.; Shi, Y.; Sun, H.; Shi, W.; Guo, F. Fabrication of Bi4Ti3O12/ZnIn2S4 S-scheme heterojunction for achieving efficient photocatalytic hydrogen production. J. Alloys Compd. 2023, 930, 167450. [Google Scholar] [CrossRef]
  80. Shi, W.; Guo, F.; Li, M.; Shi, Y.; Shi, M.; Yan, C. Constructing 3D sub-micrometer CoO octahedrons packed with layered MoS2 shell for boosting photocatalytic overall water splitting activity. Appl. Surf. Sci. 2019, 473, 928–933. [Google Scholar] [CrossRef]
  81. Guo, F.; Huang, X.; Chen, Z.; Sun, H.; Chen, L. Anchoring CoP nanoparticles on the octahedral CoO by self-phosphating for enhanced photocatalytic overall water splitting activity under visible light. Chin. J. Chem. Eng. 2021, 40, 114–123. [Google Scholar] [CrossRef]
  82. Shi, W.; Li, M.; Huang, X.; Ren, H.; Yan, C.; Guo, F. Facile synthesis of 2D/2D Co3(PO4)2/g-C3N4 heterojunction for highly photocatalytic overall water splitting under visible light. Chem. Eng. J. 2020, 382, 122960. [Google Scholar] [CrossRef]
  83. Chen, Z.; Guo, F.; Sun, H.; Shi, Y.; Shi, W. Well-designed three-dimensional hierarchical hollow tubular g-C3N4/ZnIn2S4 nanosheets heterostructure for achieving efficient visible-light photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2022, 607, 1391–1401. [Google Scholar] [CrossRef]
  84. Guo, F.; Huang, X.; Chen, Z.; Shi, Y.; Sun, H.; Cheng, X.; Shi, W.; Chen, L. Formation of unique hollow ZnSnO3@ZnIn2S4 core-shell heterojunction to boost visible-light-driven photocatalytic water splitting for hydrogen production. J. Colloid Interface Sci. 2021, 602, 889–897. [Google Scholar] [CrossRef]
  85. 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]
  86. Shi, W.; Hao, C.; Shi, Y.; Guo, F.; Tang, Y. Effect of different carbon dots positions on the transfer of photo-induced charges in type I heterojunction for significantly enhanced photocatalytic activity. Sep. Purif. Technol. 2023, 304, 122337. [Google Scholar] [CrossRef]
  87. Xu, Q.; Zhang, L.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Jaroniec, M. Direct Z-scheme photocatalysts: Principles, synthesis, and applications. Mater. Today 2018, 21, 1042–1063. [Google Scholar] [CrossRef]
  88. Xu, Z.; Shi, W.; Shi, Y.; Sun, H.; Li, L.; Guo, F.; Wen, H. Carbon dots as solid-state electron mediator and electron acceptor in S-scheme heterojunction for boosted photocatalytic hydrogen evolution. Appl. Surf. Sci. 2022, 595, 153482. [Google Scholar] [CrossRef]
  89. Wan, S.; Ou, M.; Zhong, Q.; Zhang, S.; Song, F. Construction of Z-scheme photocatalytic systems using ZnIn2S4, CoOx-loaded Bi2MoO6 and reduced graphene oxide electron mediator and its efficient nonsacrificial water splitting under visible light. Chem. Eng. J. 2017, 325, 690–699. [Google Scholar] [CrossRef]
  90. Yang, G.; Ding, H.; Chen, D.; Feng, J.; Hao, Q.; Zhu, Y. Construction of urchin-like ZnIn2S4-Au-TiO2 heterostructure with enhanced activity for photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2018, 234, 260–267. [Google Scholar] [CrossRef]
  91. Ding, Y.; Wei, D.; He, R.; Yuan, R.; Xie, T.; Li, Z. Rational design of Z-scheme PtS-ZnIn2S4/WO3-MnO2 for overall photo-catalytic water splitting under visible light. Appl. Catal. B Environ. 2019, 258, 117948. [Google Scholar] [CrossRef]
  92. Ou, M.; Li, J.; Geng, M.; Wang, J.; Wan, S.; Zhong, Q. Construction of Z-scheme Photocatalyst Containing ZnIn2S4, Co3O4-Photodeposited BiVO4 (110) Facets and rGO Electron Mediator for Overall Water Splitting into H2 and O2. Catal. Lett. 2021, 151, 2570–2582. [Google Scholar] [CrossRef]
  93. Du, X.; Zhao, T.; Xiu, Z.; Xing, Z.; Li, Z.; Pan, K.; Yang, S.; Zhou, W. BiVO4@ZnIn2S4/Ti3C2 MXene quantum dots assembly all-solid-state direct Z-Scheme photocatalysts for efficient visible-light-driven overall water splitting. Appl. Mater. Today 2020, 20, 100719. [Google Scholar] [CrossRef]
  94. Wang, Y.; Huang, W.; Guo, S.; Xin, X.; Zhang, Y.; Guo, P.; Tang, S.; Li, X. Sulfur-Deficient ZnIn2S4/Oxygen-Deficient WO3 Hybrids with Carbon Layer Bridges as a Novel Photothermal/Photocatalytic Integrated System for Z-Scheme Overall Water Splitting. Adv. Energy Mater. 2021, 11, 2102452. [Google Scholar] [CrossRef]
  95. Zuo, G.; Wang, Y.; Teo, W.L.; Xian, Q.; Zhao, Y. Direct Z-scheme TiO2–ZnIn2S4 nanoflowers for cocatalyst-free photocatalytic water splitting. Appl. Catal. B Environ. 2021, 291, 120126. [Google Scholar] [CrossRef]
  96. Zhang, J.; Zhang, Y.; Li, L.; Yan, W.; Wang, H.; Mao, W.; Cui, Y.; Li, Y.; Zhu, X. Synergizing the internal electric field and ferroelectric polarization of the BiFeO3/ZnIn2S4 Z-scheme heterojunction for photocatalytic overall water splitting. J. Mater. Chem. A 2023, 11, 434–446. [Google Scholar] [CrossRef]
  97. Yang, Y.; Sun, Z.; Liu, C.; Wang, J.; Qiu, M.; Yan, G.; Zhang, K. Boosting Photocatalytic Overall Water Splitting on Direct Z-Scheme BiOBr/ZnIn2S4 Heterostructure by Atomic-Level Interfacial Charge Transport Modulation. ACS Appl. Energy Mater. 2022, 5, 15559–15565. [Google Scholar] [CrossRef]
  98. Zuo, G.; Chen, W.; Yin, Z.; Ma, S.; Wang, Y.; Ji, Q.; Xian, Q.; Yang, S.; He, H. Covalency dominating Z-scheme perylene-dicarboximide@ZnIn2S4 organic-inorganic hybrids for overall water splitting. Chem. Eng. J. 2023, 456, 141096. [Google Scholar] [CrossRef]
  99. Zuo, G.; Ma, S.; Yin, Z.; Chen, W.; Wang, Y.; Ji, Q.; Xian, Q.; Yang, S.; He, H. Z-Scheme Modulated Charge Transfer on InVO4@ZnIn2S4 for Durable Overall Water Splitting. Small 2023, 19, e2207031. [Google Scholar] [CrossRef]
  100. Li, X.H.; Antonietti, M. Metal nanoparticles at mesoporous N-doped carbons and carbon nitrides: Functional Mott-Schottky heterojunctions for catalysis. Chem. Soc. Rev. 2014, 44, 6593–6604. [Google Scholar] [CrossRef]
  101. Chu, J.H.; Han, K.; Yu, Q.Y.; Wang, H.L.; Xi, K.; Lai, F.L.; Zhang, J.G.; Bao, Y.P. Schottky junction and multiheterostructure synergistically enhance rate performance and cycling stability. Chem. Eng. J. 2022, 430, 11. [Google Scholar] [CrossRef]
  102. Cai, X.; Zeng, Z.; Liu, Y.; Li, Z.; Gu, X.; Zhao, Y.; Mao, L.; Zhang, J. Visible-light-driven water splitting by yolk-shelled ZnIn2S4-based heterostructure without noble-metal co-catalyst and sacrificial agent. Appl. Catal. B Environ. 2021, 297, 120391. [Google Scholar] [CrossRef]
  103. Liu, X.; Zhang, J.; Xu, J.; Li, Y.; Du, Y.; Jiang, Y.; Lin, K. Hydroxyl-modified Nb4C3T(x) MXene@ZnIn2S4 sandwich structure for photocatalytic overall water splitting. J. Colloid Interface Sci. 2023, 633, 992–1001. [Google Scholar] [CrossRef]
  104. An, H.; Lv, Z.; Zhang, K.; Deng, C.; Yin, Z. Plasmonic coupling enhancement of core-shell Au@Pt assemblies on ZnIn2S4 nanosheets towards photocatalytic H2 production. Appl. Surf. Sci. 2020, 536, 147934. [Google Scholar] [CrossRef]
  105. Shi, Y.; Li, L.; Xu, Z.; Qin, X.; Cai, Y.; Zhang, W.; Shi, W.; Du, X.; Guo, F. Coupled internal electric field with hydrogen release kinetics for promoted photocatalytic hydrogen production through employing carbon coated transition metal as co-catalyst. J. Colloid Interface Sci. 2023, 630, 274–285. [Google Scholar] [CrossRef]
  106. Sun, S.; Zhang, X.; Liu, X.; Pan, L.; Zhang, X.; Zou, J. Design and Construction of Cocatalysts for Photocatalytic Water Splitting. Acta Phys.-Chim. Sin. 2020, 36, 1905007–1905000. [Google Scholar] [CrossRef]
  107. Li, H.; Xiao, J.; Vequizo, J.J.M.; Hisatomi, T.; Nakabayashi, M.; Pan, Z.; Shibata, N.; Yamakata, A.; Takata, T.; Domen, K. One-Step Excitation Overall Water Splitting over a Modified Mg-Doped BaTaO2N Photocatalyst. ACS Catal. 2022, 12, 10179–10185. [Google Scholar] [CrossRef]
  108. Ma, B.; Dang, Y.; Li, D.; Wang, X.; Lin, K.; Wang, W.; Zhou, X.; Chen, Y.; Xie, T.; Zhang, X. A Yin-Yang hybrid co-catalyst (CoOx-Mo2N) for photocatalytic overall water splitting. Appl. Catal. B Environ. 2021, 298, 120491. [Google Scholar] [CrossRef]
  109. Jing, H.; Li, H.; Yue, J.; Fan, S.; Yao, B.; Liu, S.; Fang, Z.; Wu, R.; Yang, L.; Wei, S. Synergistic effects of the Rh-S bond and spatially separated dual cocatalysts on photocatalytic overall water splitting activity of ZnIn2S4 nanosheets under visible light irradiation. Dalton Trans. 2023, 52, 2924–2927. [Google Scholar] [CrossRef]
Scheme 1. Modification strategies for ZnIn2S4-based photocatalysts in photocatalytic overall water splitting are briefly described.
Scheme 1. Modification strategies for ZnIn2S4-based photocatalysts in photocatalytic overall water splitting are briefly described.
Catalysts 13 00967 sch001
Scheme 2. Schematic diagram of ZnIn2S4-based photocatalyst photocatalytic overall water splitting.
Scheme 2. Schematic diagram of ZnIn2S4-based photocatalyst photocatalytic overall water splitting.
Catalysts 13 00967 sch002
Figure 1. Crystal structure of cubic (a), hexagonal (b), rhombic (c) ZnIn2S4 [37].
Figure 1. Crystal structure of cubic (a), hexagonal (b), rhombic (c) ZnIn2S4 [37].
Catalysts 13 00967 g001
Figure 2. (a) Schematic illustration of the dual-defect (Ag dopants and nanoholes) configuration established on ZnIn2S4 monolayers and (b) comparison of the H2 and O2 evolution activities of ZnIn2S4, Ag-ZnIn2S4, and Cu-ZnIn2S4 in photocatalytic water-splitting reaction under visible-light illumination (300 W xenon lamp, λ > 420 nm) [69]. The differential charge density maps for pristine-ZIS atomic layers (c) and dZni-ZIS atomic layers (d), and (e) X-ray diffraction (XRD) patterns of pristine-ZIS and dZni-ZIS [70].
Figure 2. (a) Schematic illustration of the dual-defect (Ag dopants and nanoholes) configuration established on ZnIn2S4 monolayers and (b) comparison of the H2 and O2 evolution activities of ZnIn2S4, Ag-ZnIn2S4, and Cu-ZnIn2S4 in photocatalytic water-splitting reaction under visible-light illumination (300 W xenon lamp, λ > 420 nm) [69]. The differential charge density maps for pristine-ZIS atomic layers (c) and dZni-ZIS atomic layers (d), and (e) X-ray diffraction (XRD) patterns of pristine-ZIS and dZni-ZIS [70].
Catalysts 13 00967 g002
Figure 3. (a) X-ray diffraction (XRD) patterns of ZnIn2S4 and ZnIn2S4-800, (b) Electron paramagnetic resonance (EPR) results of ZnIn2S4 and ZnIn2S4-800, and (c) Ultraviolet–visible (UV−vis) light absorption spectra and (d) Band-gap energies of ZnIn2S4 and ZnIn2S4-800 [75].
Figure 3. (a) X-ray diffraction (XRD) patterns of ZnIn2S4 and ZnIn2S4-800, (b) Electron paramagnetic resonance (EPR) results of ZnIn2S4 and ZnIn2S4-800, and (c) Ultraviolet–visible (UV−vis) light absorption spectra and (d) Band-gap energies of ZnIn2S4 and ZnIn2S4-800 [75].
Catalysts 13 00967 g003
Figure 4. (a) Z-scheme photocatalytic water splitting systems consisting of H2-evolving photocatalyst (Pt/ZnIn2S4), O2-evolving photocatalyst (CoOx/BMO), and electron medium (RGO) [89]. (b) Schematic illustration of the fabrication process of the ZnIn2S4-Au-TiO2 photocatalyst: Au-TiO2 is prepared by a chemical-deposition process and ZnIn2S4-Au-TiO2 is synthesized by the solvothermal process. (c) UV-Vis diffuse reflectance spectra of TiO2, Au-TiO2, ZnIn2S4-Au-TiO2, and ZnIn2S4 [90]. (d) Proposed mechanism for photocatalytic overall water splitting over PtS-ZnIn2S4/WO3-MnO2 nanocomposites under visible light [91].
Figure 4. (a) Z-scheme photocatalytic water splitting systems consisting of H2-evolving photocatalyst (Pt/ZnIn2S4), O2-evolving photocatalyst (CoOx/BMO), and electron medium (RGO) [89]. (b) Schematic illustration of the fabrication process of the ZnIn2S4-Au-TiO2 photocatalyst: Au-TiO2 is prepared by a chemical-deposition process and ZnIn2S4-Au-TiO2 is synthesized by the solvothermal process. (c) UV-Vis diffuse reflectance spectra of TiO2, Au-TiO2, ZnIn2S4-Au-TiO2, and ZnIn2S4 [90]. (d) Proposed mechanism for photocatalytic overall water splitting over PtS-ZnIn2S4/WO3-MnO2 nanocomposites under visible light [91].
Catalysts 13 00967 g004
Figure 5. (a) Schematic illustration for the formation of BV@ZIS/TC QDs assembly, (b) Schematic illustration of band structure and electron–hole transfer mechanism for BV@ZIS/TC QDs [93].
Figure 5. (a) Schematic illustration for the formation of BV@ZIS/TC QDs assembly, (b) Schematic illustration of band structure and electron–hole transfer mechanism for BV@ZIS/TC QDs [93].
Catalysts 13 00967 g005
Figure 6. (a) Schematic illustration for the synthetic process of TNZIS, (b) Schematic illustration for the reaction mechanism of photocatalytic water splitting over direct Z-scheme TNZIS heterojunction, (c) Calculated average potential profile along the Z axis of TiO2 and ZnIn2S4. (d) Difference in charge density isosurface at the interface between TiO2 and ZnIn2S4. Ti: blue, O: red, Zn: silver, In: purple, and S: yellow sticks. Blue and yellow areas indicate the loss and accumulation of electrons, respectively. * Total net charge is derived by the sum of Bader atomic charges on the TiO2 layer. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article) [95].
Figure 6. (a) Schematic illustration for the synthetic process of TNZIS, (b) Schematic illustration for the reaction mechanism of photocatalytic water splitting over direct Z-scheme TNZIS heterojunction, (c) Calculated average potential profile along the Z axis of TiO2 and ZnIn2S4. (d) Difference in charge density isosurface at the interface between TiO2 and ZnIn2S4. Ti: blue, O: red, Zn: silver, In: purple, and S: yellow sticks. Blue and yellow areas indicate the loss and accumulation of electrons, respectively. * Total net charge is derived by the sum of Bader atomic charges on the TiO2 layer. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article) [95].
Catalysts 13 00967 g006
Figure 7. (a) Band structures of the BiOBr/ZnIn2S4 (BOB/ZIS) direct Z-scheme heterojunction; (b) X-ray photoelectron spectroscopy (XPS) spectra O 1s; (c) FTIR spectra of composite samples [97]. (d) Schematic illustration of HPZ synthesis, (e) Z axis potential profile of HC-PDI and ZnIn2S4. (f) Differential charge density between HC-PDI and ZnIn2S4. Zn: silver, In: purple, S: yellow, C: brown, O: red, and H: white spheres [98].
Figure 7. (a) Band structures of the BiOBr/ZnIn2S4 (BOB/ZIS) direct Z-scheme heterojunction; (b) X-ray photoelectron spectroscopy (XPS) spectra O 1s; (c) FTIR spectra of composite samples [97]. (d) Schematic illustration of HPZ synthesis, (e) Z axis potential profile of HC-PDI and ZnIn2S4. (f) Differential charge density between HC-PDI and ZnIn2S4. Zn: silver, In: purple, S: yellow, C: brown, O: red, and H: white spheres [98].
Catalysts 13 00967 g007
Figure 8. (a) Schematic illustration of the synthetic process of NiCo2S4/ZnIn2S4/Co3O4 heterostructure, and band alignment of ZnIn2S4 and NiCo2S4 before (b) and after (c) their connection [102].
Figure 8. (a) Schematic illustration of the synthetic process of NiCo2S4/ZnIn2S4/Co3O4 heterostructure, and band alignment of ZnIn2S4 and NiCo2S4 before (b) and after (c) their connection [102].
Catalysts 13 00967 g008
Figure 9. (a) Schematic illustration of the synthetic process of Nb4C3Tx MXene@ ZnIn2S4-OH sandwich composite, (b) Scanning kelvin probe (SKP) maps of (c) Nb4C3Tx MXene and ZnIn2S4 [103].
Figure 9. (a) Schematic illustration of the synthetic process of Nb4C3Tx MXene@ ZnIn2S4-OH sandwich composite, (b) Scanning kelvin probe (SKP) maps of (c) Nb4C3Tx MXene and ZnIn2S4 [103].
Catalysts 13 00967 g009
Table 1. Summary of different morphologies and synthesis methods for hexagonal phase ZnIn2S4 photocatalysts.
Table 1. Summary of different morphologies and synthesis methods for hexagonal phase ZnIn2S4 photocatalysts.
TypeMorphologyPhotocatalystSynthetic MethodSulfur SourceSolventLight SourceApplicationRef
0Dquantum dotsZnIn2S4solvothermalsulfur powderoctadecene500 W Xe lampdegradation[40]
1DnanowiresZnIn2S4wet-chemicalthioacetamide (TAA)H2O500 W Xe lampdegradation[41]
1DnanotubesZnIn2S4wet-chemicalTAAH2O500 W Xe lampdegradation[41]
2Dultrathin nanosheetVs-M-ZnIn2S4lithium intercalationTAAN,N-Dimethylformamide, ethylene glycol300 W Xe lamphydrogen generation[44]
3Dpersimmon-like shapeZnIn2S4solvothermalCS2tetrahydrofuran (THF)300 W Xe lamphydrogen generation[45]
3Dporous ZnIn2S4
submicrospheres
ZnIn2S4microwave-solvothermalexcess thioureaethylene glycol300 W tungsten–halogendegradation[46]
3Dpeony-flower-likeZnIn2S4solventhermaldioctyldithiocarbamic acid sodium (OTC)CH3OH300 W tungsten-halogendegradation[47]
3Drose-flower-like microsphereZnIn2S4hydrothermalthioureaH2O, diethyl amine (DEA)300 W Xe lamphydrogen generation[49]
3Dhollow marigold-like flowersZnIn2S4hydrothermalthioureaH2O, polyvinyl pyrrolidone (PVP)300 W Xe lamphydrogen generation[49]
3Dporous microspheresZnIn2S4microwave-sol
vothermal
excessive TAAH2O500 W tungsten–halogen lampdegradation[50]
3Dhollow StructureZnIn2S4hydrothermalglutathione (GSH)H2O300 W Xe lamphydrogen generation[51]
Table 2. Summary of reports on photocatalytic overall water splitting of ZnIn2S4-based catalysts.
Table 2. Summary of reports on photocatalytic overall water splitting of ZnIn2S4-based catalysts.
PhotocatalystCocatalystReaction SystemsLight SourceH2 (μmol
g−1 h−1)
O2 (μmol
g−1 h−1)
AQERef.
Ag-ZnIn2S4/12 mg (100 mL H2O)300 W Xe lamp
(λ > 420 nm)
56.629.10.70% (405 nm)
0.57% (420 nm)
0.20% (450 nm)
[69]
dZni-ZnIn2S4/50 mg (120 mL H2O)300 W Xe lamp
(λ > 420 nm)
42.819.11.51% (420 nm)[70]
Al-ZnIn2S4/50 mg (100 mL H2O)300 W Xe lamp
(λ ≥ 420 nm)
77.235.31.61% (420 nm)[71]
Sv-ZnIn2S4-O (ZnIn2S4-350 °C-4 h)Pt/Cr50 mg (100 mL H2O)300 W Xe lamp
(λ ≥ 420 nm)
270.2130.00.21% (420 nm)[72]
Sv-ZnIn2S4 (ZnIn2S4-800)Pt/Cr50 mg (100 mL H2O)300 W Xe lamp
(λ ≥ 420 nm)
68.031.00.041% (420 nm)
0.016% (450 nm)
0.004% (500 nm)
[75]
ZnIn2S4/RGO/BMOPt/CoOx100 mg (100 mL H2O)200 W Xe lamp
(λ > 420 nm)
31.415.8/[89]
ZnIn2S4-Au-TiO2/50 mg (100 mL H2O)300 W Xe lamp186.366.3/[90]
Pt-ZnIn2S4/RGO/Co3O4-BiVO4(110)Pt/Co3O450 mg (100 mL H2O)300 W Xe lamp(λ > 420 nm)24.511.9/[92]
PtS-ZnIn2S4/WO3-MnO2Pt/MnO250 mg (100 mL H2O)300 W Xe lamp
(λ > 420 nm)
38.815.7/[91]
BiVO4@ZnIn2S4/Ti3C2 MXene QDsTi3C2 MXene QDs60 mg (H2O)300 W Xe lamp
(λ > 400 nm)
102.6750.832.40% (410 nm)
2.90% (460 nm)
1.40% (510 nm)
0.20% (560 nm)
[93]
ZIS-WO/C-wood (Sv-ZnIn2S4-Ov-WO3/C-wood)Pt/CoOxfloated at the water–air interface300 W Xe lamp
(AM 1.5G)
169.282.5/[94]
TiO2-ZnIn2S4/20 mg (50 mL H2O)300 W Xe lamp214.981.7/[95]
BiFeO3/ZnIn2S4/12 mg (100 mL H2O)300 W Xe lamp
(λ > 420 nm)
87.342.31.12% (420 nm)[96]
BiOBr/ZnIn2S4Pt100 mg (100 mL H2O)300 W Xe lamp
(λ > 420 nm)
628304/[97]
HC-PDI@ZnIn2S4 OIHs/5 mg (50 mL H2O)300 W Xe lamp
(λ ≥ 400 nm)
275.4138.416.14% (400 nm)[98]
InVO4@ZnIn2S4/5 mg (50 mL H2O)300 W Xe lamp153.376.924.28% (365 nm)
19.31% (380 nm)
15.29% (400 nm)
9.75% (420 nm)
6.93% (460 nm)
[99]
NiCo2S4/ZnIn2S4/Co3O4NiCo2S4/Co3O410 mg (15 mL H2O)
/103.326.7/[102]
Nb4C3Tx MXene@ZnIn2S4-OHNb4C3Tx MXene/OH20 mg (100 mL H2O)300 W Xe lamp
(λ > 420 nm)
53.826.7/[103]
ZnIn2S4-Rh-CrRh-Cr50 mg (100 mL H2O)300 W Xe lamp
(AM 1.5G)
118580.084% (420 nm)
0.028% (450 nm)
0.017% (500 nm)
[109]
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

Yan, Y.; Chen, Z.; Cheng, X.; Shi, W. Research Progress of ZnIn2S4-Based Catalysts for Photocatalytic Overall Water Splitting. Catalysts 2023, 13, 967. https://doi.org/10.3390/catal13060967

AMA Style

Yan Y, Chen Z, Cheng X, Shi W. Research Progress of ZnIn2S4-Based Catalysts for Photocatalytic Overall Water Splitting. Catalysts. 2023; 13(6):967. https://doi.org/10.3390/catal13060967

Chicago/Turabian Style

Yan, Yujie, Zhouze Chen, Xiaofang Cheng, and Weilong Shi. 2023. "Research Progress of ZnIn2S4-Based Catalysts for Photocatalytic Overall Water Splitting" Catalysts 13, no. 6: 967. https://doi.org/10.3390/catal13060967

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop