Next Article in Journal
Plasma Controlled Growth Dynamics and Electrical Properties of Ag Nanofilms via RF Magnetron Sputtering
Previous Article in Journal
Comment on Rasool et al. Spectral Relaxation Methodology for Chemical and Bioconvection Processes for Cross Nanofluid Flowing Around an Oblique Cylinder with a Slanted Magnetic Field Effect. Coatings 2022, 12, 1560
Previous Article in Special Issue
Fabrication and Mechanism Investigation of High-Porosity Micro-Arc Oxidation Functional Coating on Aluminum Foam Substrate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Defect Engineering of ZnIn2S4 Photocatalysts for Enhanced Hydrogen Evolution Reaction

1
Shandong Key Laboratory of Water Pollution Control and Resource Reuse, Shandong Key Laboratory of Environmental Processes and Health, School of Environmental Science and Engineering, Shandong University, Qingdao 266237, China
2
Weihai Research Institute of Industrial Technology of Shandong University, Shandong University, Weihai 264209, China
3
State Key Laboratory of Microbial Technology, Microbial Technology Institute, Shandong University, Qingdao 266237, China
*
Authors to whom correspondence should be addressed.
Coatings 2025, 15(9), 1061; https://doi.org/10.3390/coatings15091061
Submission received: 31 July 2025 / Revised: 28 August 2025 / Accepted: 9 September 2025 / Published: 10 September 2025

Abstract

ZnIn2S4, a visible-light-responsive layered sulfide photocatalyst with a suitable bandgap (~2.4 eV), exhibits considerable potential for the photocatalytic hydrogen evolution reaction (PHER) due to its low toxicity, excellent stability, and appropriate band alignment. Nevertheless, its practical deployment is limited by inherent issues such as rapid charge carrier recombination, scarce surface-active sites, and slow oxidation kinetics. Defect engineering strategies—including sulfur, zinc, and indium vacancies, as well as heteroatom doping—have been developed to mitigate these shortcomings. This review not only summarizes recent advances in these strategies but also elucidates the fundamental physicochemical mechanisms behind the enhanced photocatalytic performance. A systematic quantitative evaluation is presented, highlighting improvements in critical performance metrics such as hydrogen evolution rate, light absorption range, apparent quantum yield (AQY), and charge separation efficiency. Furthermore, the review offers a critical perspective on the current state of defect-engineered ZnIn2S4 systems. Promising future research pathways are outlined, with emphasis on atomic-precision synthesis and operando characterization techniques. Finally, we discuss persistent challenges in the field, including reproducibility in synthesis, long-term operational stability, and scalability toward industrial hydrogen production.

1. Introduction

Mitigating the intensifying climate crisis necessitates a global transition towards carbon neutrality, demanding transformative technologies for clean energy production and storage [1,2,3,4,5]. Hydrogen, envisioned by the International Energy Agency (IEA) to constitute approximately 13% of the world’s final energy consumption by 2050, emerges as a pivotal clean energy carrier. However, the current hydrogen landscape is dominated by production methods reliant on fossil fuel reforming, which incur significant carbon dioxide emissions, which starkly contradicts the imperative for decarbonization. Consequently, the development of sustainable and carbon-neutral hydrogen evolution reaction HER technologies is critically urgent [6,7,8,9]. Among the promising solutions, photocatalytic hydrogen evolution reaction (PHER) through water splitting stands out [10,11,12,13]. This approach harnesses abundant solar energy to directly dissociate water molecules, generating storable hydrogen without the associated carbon emissions [14,15,16,17]. By mimicking natural photosynthesis, photocatalysis offers a potentially scalable pathway to convert solar energy directly into chemical fuels, representing a cornerstone strategy for a sustainable hydrogen economy [18,19,20,21].
Among visible-light photocatalysts, hexagonal ZnIn2S4 (ZIS, Eg ≈ 2.4 eV) exhibits compelling advantages [22,23,24,25]. Its hexagonal phase features advantageous [S–Zn–In–S] layers that facilitate charge transport and modification, combined with optimal band positions, good chemical stability, and an eco-friendly composition [26,27,28,29]. However, three key limitations impede its practical application as shown in Figure 1 [30,31,32,33]:
(i)
Severe charge recombination: Due to inter-layer potential barriers (>0.5 eV) and deep-trap states (0.8–1.4 eV below the conduction band minimum) induced by sulfur vacancies (VS), which function as Auger-mediated recombination centers. These constraints confine carrier transport primarily to out-of-plane pathways, culminating in >80% bulk recombination within 5 ns rapid carrier recombination due to inter-layer barriers and inherent defects [15,34,35].
(ii)
Insufficient HER kinetics: Arising from low active-site density (<5% S-atom utilization). Density functional theory (DFT) calculations reveal that this originates from suboptimal orbital hybridization: the d-band center of surface In atoms resides at −2.8 eV below EF, weakening S 3p–H 1s coupling. Consequently, the Gibbs free energy of hydrogen adsorption (ΔGH) deviates substantially from thermoneutrality (ΔGH ≈ 0.35 eV), placing ZIS on the weak-adsorption branch of the HER [36,37,38].
(iii)
Photocorrosion instability: Mediated by VS sites, which initiate an autocatalytic degradation cycle. Under illumination, VS-derived trap states (0.8 eV below CBM) accumulate photogenerated holes, oxidizing adjacent S atoms and releasing H2S, thereby inducing irreversible structural collapse [39,40,41,42].
Recent surface engineering strategies, such as defect engineering, heterogeneous structure construction, surface functionalization, morphology and facet control, have been effectively addressing these issues. For instance, surface vacancies (such as sulfur, zinc, or indium vacancies) induce trap-dominated recombination by optimizing conduction band alignment or reconstructing coordination environments to suppress sulfur oxidation thermodynamics [43,44]; metal or non-metal decoration optimizes hydrogen adsorption (ΔGH = −0.03 eV) and suppresses interface recombination, and single-atom metal decoration will amplify active-site density [45]; heterojunction construction achieves spatial charge separation through built-in fields [46,47,48] and also the surface protective layer to suppress sulfur oxidation [49,50]. It should be noted that the defect engineering provides a precise solution to overcome the three issues listed in Figure 1, through atomic-scale interface manipulation (sulfur, zinc, or indium vacancies; heteroatoms doping), enhancing charge separation efficiency, optimizing active-site density, and suppressing photocorrosion via protective coatings or defect passivation [51,52].
Herein, this comprehensive review critically discusses defect engineering strategies specifically tailored for ZIS-based photocatalysts to advance the photocatalytic hydrogen evolution reaction (PHER). We systematically examine three major modification approaches, including sulfur, zinc, or indium vacancies, as well as heteroatom doping. Beyond summarizing these strategies, the review elucidates the underlying physicochemical mechanisms responsible for enhanced performance. A rigorous quantitative assessment benchmarks the improvements in key metrics, such as the PHER rate, light absorption, apparent quantum yield (AQY), and carrier separation efficiency. Importantly, the review adopts a critical perspective toward these surface-engineered ZIS photocatalytic systems. Furthermore, it highlights promising future research directions with an emphasis on atomic-precision synthesis and operando diagnostics. Finally, we address ongoing challenges related to synthesis reproducibility, long-term operational stability, and process integration for large-scale HER.

2. Classification and Mechanisms of Defect Engineering Strategies

2.1. Sulfur Vacancies (VS)

Sulfur vacancy (VS) engineering represents an effective approach to enhancing optical, electrical, and surface properties, thereby improving PHER performances [53]. As a typical anion defect in ZIS, VS induce lattice distortion by breaking the local coordination balance of S atoms, thus reconstructing the electronic structure of the material [54]. Introducing VS into ZIS is a common defect engineering strategy aimed at improving its photocatalytic performance (such as promoting the separation of photoinduced charge carriers, enhancing adsorption capacity, etc.). VS can serve as active sites or alter the material’s electronic structure. VS serve as efficient electron traps by introducing mid-gap defect states in ZIS, which significantly enhance light absorption and prolong photogenerated carrier lifetimes through suppressed recombination. Also, the VS site could act as active sites for the absorption of H2O molecules and accelerate the HER. As shown in Figure 2, the formation of VS could be obtained by creating a sulfur-deficient chemical environment or applying energy (heat, plasma, radiation) to facilitate the escape of sulfur atoms, which mainly rely on the following methods: high-temperature calcination, hydrothermal/solvothermal reduction, plasma treatment technology, and other high-energy application methods. Although the introduction of VS in ZIS could enhance the PHER by creating defect energy levels, the control and modulation of VS are very important to PHER performances. Moderate concentrations of VS enhance charge separation by trapping photogenerated electrons/holes, thereby suppressing recombination and boosting the HER. However, excessive VS formation may establish recombination centers that degrade photocatalytic efficiency. Therefore, there is a critical need for precise VS modulation.

2.1.1. High-Temperature Calcination Method

The calcination atmospheres (such as air, N2, Ar, H2/Ar, and so on) can create a sulfur-deficient chemical environment to achieve selective desulfurization at 200–800 °C, which facilitates the escape of sulfur atoms for the formation of VS [55,56,57]. And the gradient control for the density of VS can be realized by adjusting the calcination temperature and treated time. This method offers the advantages of a clear underlying principle; relatively precise control over vacancy concentration via temperature, holding duration, and gas concentration parameters; and scalability for large-scale treatment. As shown in Figure 3a–h, Jing et al. determined to conduct calcination experiments at temperatures below 500 °C by comparing the crystal phase conditions at 350 °C and 500 °C for different durations for the modulation of VS [56]. ZIS engineered with VS with the highest HER performances were fabricated through controlled calcination at 350 °C for 4h. The optimized light absorption characteristics and extending charge carrier lifetimes were obtained. Consequently, the ZIS-350-4h photocatalyst achieved an apparent quantum efficiency (AQE) of 0.21% at 420 nm and a solar-to-hydrogen (STH) conversion efficiency of 0.035%—representing the highest reported performance for pristine ZIS to date. As shown in Figure 3i–p, they also prepared rhombohedral ZIS nanosheets enriched with VS under a N2 atmosphere at 400 °C and 800 °C, ZIS enriched with VS obtained at 800 °C with Pt/Cr cocatalysts presents a HER rate of 3.40 μmol·h−1 [55]. An appropriate amount of VS can improve the overall water splitting performance; the STH conversion efficiency of sulfur vacancy-enriched rhombohedral ZIS reaches 0.021%, which is the highest reported efficiency for single-phase ZIS. As shown in Figure 3k, the shift in S 2p3/2 (~161.4 eV) and S 2p1/2 (~162.6 eV) peaks in S 2p XPS spectra corresponded to the presence of low-valence sulfur or defect-associated sulfur states. Additionally, a decreased S/(Zn + In) atomic ratio serves as further evidence of VS. Electron paramagnetic resonance (EPR) directly captures the characteristic signals of unpaired electrons localized at VS, and its intensity is linearly related to the vacancy density (see Figure 3d,l). As shown in Figure 3m, ZIS-800 shows a red-shift in the absorption edge in the UV-vis spectra, which is generally associated with various defects, further confirming the presence of rich VS and are consistent with the EPR results. Moreover, ZIS-800 has a lower PL intensity, indicating a higher separation efficiency of photogenerated electron–hole pairs. The lifetime of ZIS-800 (1.902 ns) is higher than that of ZIS (1.506 ns) in TRPL spectra (Figure 3n), effectively avoiding carrier recombination, and VS as active sites can significantly improve the HER rate. Similarly, Huang et al. controlled ZIS that engineered VS via calcination under a N2 atmosphere for two hours at a much lower temperature (300 °C) for piezo-catalytic HER; a high HER rate of approximately 3164.67 µmo1∙g−1∙h−1 under ultrasound was obtained with the optimal defect level [58].

2.1.2. Hydrothermal/Solvothermal Reduction Method

During the hydrothermal/solvothermal reaction, sulfur vacancy formation is facilitated by controlling reaction conditions (temperature, time, and pH) [59,60]. This reduces the stability of the sulfur source or inhibits its complete reaction, allowing VS to form naturally during crystal growth. Alternatively, different sulfur sources are used (e.g., thioacetamide (TAA) decomposes more readily than Na2S and thiourea) or complexing agents are added to modulate reaction kinetics during the hydrothermal/solvothermal reaction to obtain VS. The approach provides relative process simplicity while enabling simultaneous control over both material morphology and sulfur vacancy concentration. Specifically, the vacancy concentration can be precisely tuned by modifying key parameters such as the reductant amount, reaction temperature, and time. This is proved by Xu et al., who selected an appropriate sulfur source (TAA) and optimized the hydro/solvothermal conditions to generate a certain level of VS [61]. As shown in Figure 4, they investigated various sulfur vacancy levels in pristine ZIS (denoted as ZIS-X; and P, M, and R represent poor, moderate, and rich contents of VS, respectively), proved in the EPR spectra. ZIS-M with a moderate level of sulfur vacancies demonstrated the highest PHER performance of around 25 mmol·g−1. ZIS-M, with the calculated theoretical work functions of 4.14 eV, possessed the most favorable hydrogen adsorption-free energy, resulting in superior HER performances. Interestingly, Ren et al. fabricated VS-ZIS with efficient PHER performances (5437 μmol−1·h−1∙g−1) via a one-step solvothermal process using ethylene glycol (EG); the addition of EG plays a key role in producing defects in the VS during the solvothermal reaction [62].
In addition, a moderate amount of mild reducing agent, such as hydrazine hydrate (N2H4·H2O), sodium borohydride (NaBH4) or sodium hypophosphite (NaH2PO2) can be added; these reductants reduce part of the sulfur ions (S2−), causing them to escape as hydrogen sulfide (H2S) and leaving VS during the reaction process. It should be noted here that these agents require extreme care as they are highly reducing, toxic, and flammable at elevated temperatures; the experiments should be operated strictly in a fume hood and with precise controlled quantities. ZIS engineered by VS was prepared via a N2H4·H2O-assisted hydrothermal method. Typically, 100 mg as-synthesized ZIS was dispersed into 20 mL of deionized water for 1 h; then, 5 mL of N2H4·H2O was added into the mixing solution and stirred for another 30 min. After that, the mixture was transferred to a 50 mL stainless steel autoclave and maintained at 240 °C oven for 5 h. Finally, the precipitate was separated by centrifugation and washed with deionized water several times, then dried at 60 °C for 10 h. The obtained light-yellow powder was labeled as ZIS engineered by VS [63]. Reducing agents (e.g., NaBH4) drive in situ formation of VS during crystal growth. Li et al. utilized this method to anchor Fe-Ni2P cocatalysts onto ZIS nanosheets rich with VS [64]. The sites of VS acted as nucleation centers for atomic-level interfacial bonding, facilitating electron transfer and yielding a HER rate of 15.6 mmol·g−1·h−1.

2.1.3. Plasma Treatment Technology

The generation of VS in ZIS via plasma treatment operates through kinetic energy transfer and reactive species etching. Plasma-generated ions (e.g., Ar+, H+) or electrons (10–100 eV kinetic energy) collide with surface S atoms. Energy transfer exceeds the S-Zn/In bond dissociation energy (~2–3 eV), ejecting S atoms via momentum transfer, leading to the formation of VS. Liu et al. reported ZIS nanosheets with VS (ZIS-VS) synthesized through Ar plasma treatment, and its defect concentration was modulated by varying the treatment times, which were confirmed by EPR spectra [65]. As shown in Figure 5a–g, Ar plasma-treated ZIS with optimized sulfur vacancies (radio frequency discharge: 70 W, 13.56 MHz; duration: 5 min) achieves a visible-light PHER rate of 261.9 μmol·h−1, representing a 2.06-fold enhancement over pristine ZIS. The AQE reaches 51.2% at 420 nm of monochromatic irradiation. H2O-adsorption FTIR spectroscopy confirms the enhanced adsorption of water molecules at vacancy sites. Furthermore, density functional theory (DFT) calculations elucidate the altered HER kinetics in alkaline media between VS-ZIS and pristine ZIS. These results demonstrate that Ar plasma treatment effectively engineers surface sulfur vacancies on ZIS nanosheets to boost the PHER efficiency. Jiang et al. carried out a comparative study on VS-ZIS obtained through two facile approaches, plasma etching and annealing, for enhancing the photocatalytic performance [66]. The defect formation energies for atoms at various sites within a single unit cell of ZIS were calculated and are presented in Figure 5i. From the density of states (DOS) of both ZIS and VS-ZIS in Figure 5j,k, it is evident that the valence band maximum (VBM) of the ZIS microsphere is primarily composed of S 2p and In 2s orbitals, while the conduction band minimum (CBM) is dominated by S 2p orbitals. In contrast to pristine ZIS, a deep defect level emerges within the bandgap of VS-ZIS, contributed mainly by S 2p, S 2s, and In 2s orbitals. This localized state lies above the Fermi level and represents an occupied electron state. It is widely recognized that such states can enhance electron–hole separation by trapping and exciting electrons; however, they may also act as recombination centers due to relatively low hole mobility, which could potentially reduce photocatalytic activity. Therefore, Ar plasma-etched VS-ZIS achieved an HER of 706 μmol·g−1·h−1, which was 5 and 1.2 times higher than that of pure ZIS and annealed ZIS, respectively (see Figure 5l,m), due to optimized surface-localized VS, whereas bulk vacancies from annealing acted as recombination centers.
In H2/Ar plasma, reactive H radicals reduce surface S2− resulting in sulfur etching and leaving VS. For instance, ZIS nanosheet arrays were modified by H2-Ar plasma to introduce VS; EPR spectra unambiguously confirmed the formation of VS. By varying plasma power (20–100 W), the modulation of the concentration of VS was controlled. This defect engineering significantly enhanced photoelectrochemical performance: plasma treatment at 60 W yielded a photocurrent density of 0.3 mA cm−2 at 0.3 V of VS. RHE—approximately double that of untreated samples—with a corresponding HER rate of 2.74 μmol∙cm−2∙h−1 versus 0.96 μmol∙cm−2∙h−1 for pristine ZIS material. These performance enhancements validate EPR as a reliable quantitative indicator of vacancy states [67].

2.1.4. Other Methods

Other high-energy methods (such as ball milling and laser) were also used to generate VS for ZIS. The VS were formed predominantly through mechanochemical reactions and lattice defect engineering during high-energy ball milling (HEBM) [68]. Intense mechanical collisions and friction transfer substantial energy to the ZIS material, facilitating sulfur diffusion and migration from the lattice to create VS. Concurrently, HEBM induces structural modifications including reduced crystallite size and increased lattice distortion, which collectively regulate the distribution density of VS. The thermal hotspots formed during mechanical energy conversion also accelerate surface property modifications, creating additional catalytic active sites beneficial for photocatalytic reactions. Beyond VS, HEBM introduces co-existing defects such as Zn/In vacancies and interstitial atoms. Daria Roda et al. proposed a new method for preparing ZIS thin films using pulsed laser deposition (PLD) [69]. Future research should focus on improving the durability of ZIS through methods such as protective coatings and doping strategies, so as to mitigate degradation and maintain a high photoelectrochemical (PEC) performance.
The advancement of the engineering of VS in photocatalysts faces three critical challenges: (1) regarding stability limitations, high-concentration sites of VS are prone to oxidation by reactive oxygen species; (2) precision control demands, as existing methods lack an atomic-layer spatial resolution for the distribution of VS, requiring hybrid techniques such as ALD-EBID for site-specific vacancy creation; and (3) cross-scale mechanism gaps. The intrinsic charge transport properties and the dynamic evolution of VS under reaction conditions fail to form a coherent mechanism chain spanning atomic, mesoscopic, and macroscopic scales. As the core regulatory unit of ZIS surface engineering, VS break through the bottleneck of the PHER through the trinity mechanism of “electron capture-bandgap narrowing-adsorption optimization”. In the future, it is necessary to integrate atomic manufacturing, in situ diagnosis and machine learning to achieve a precise editing of the concentration/position of VS and promote the practical application of solar HER [58].

2.2. Zinc or Indium Vacancies

In the PHER, zinc (Zn) or indium (In) vacancies in the surface engineering of ZIS significantly enhance catalytic performances by regulating the electronic structure, optimizing the active sites, and improving the carrier separation efficiency [70,71]. Xu et al. reported that ZIS nanosheets with controlled Zn vacancies (VZn) had an efficient PHER rate of 212.0 μmol under 4h light irradiation, which is 1.9 times that of pristine ZIS [72]. As shown in Figure 6a, VZn-ZIS was synthesized by calcination at 200 °C for 20 min in air. The composition of the CBM for both ZIS and VZn-ZIS, derived from the DOS calculations in Figure 6b, reveals a significantly enhanced DOS near the Fermi level. This suggests that VZn can introduce new defect states below the CBM and promote the accumulation of photogenerated electrons in the conduction band of ZIS. Furthermore, the partial charge density distributions of ZIS and VZn-ZIS visually demonstrate that the charge density near the CBM of pristine ZIS is primarily localized around the In atoms and adjacent S atoms, consistent with the DOS results. Notably, a pronounced increase in charge density is observed on S atoms near the VZn sites, confirming that Zn vacancies facilitate the enrichment of photoexcited electrons on these sulfur atoms and promote efficient charge transfer. In ZIS, the S atoms surrounding Zn exhibit strong electron localization, leading to a high adsorption but poor desorption capability for H*. In contrast, VZn-ZIS shows a more uniform redistribution of electron density due to interactions among neighboring S atoms, resulting in a balanced H* adsorption and desorption behavior. The terminal S site adjacent to the VZn exhibits an excellent hydrogen adsorption-free energy (ΔGH*) of 0.19 eV, indicating optimal adsorption/desorption properties that are highly beneficial for the PHER. As present in Figure 6e, the defect states of charge density enhanced by VZn near the ZIS’s CB promote the migration of photogenerated carriers around Zn vacancies. In addition, the VZn can redistribute the charge density of the surrounding S atoms, resulting in high-efficiency surface HER sites. Also, the VZn-ZIS could be obtained by the treatments of 5% HF solution as shown in Figure 6f–i; the introduction of VZn does not affect the lattice structure of ZIS but causes progressively weaker visible-light absorption and an increase in the bandgap values, which is beneficial for the separation of photogenerated carriers [73]. Among these samples, VZn-ZIS24 exhibited the highest HER rate of 4.18 mmol∙g−1∙h−1, which was 1.42 times that of pristine ZIS. The photocurrent data of VZn-ZIS proved the enhanced carrier separation and charge transfer that resulted from the introduction of VZn. Additionally, VZn-ZIS exhibits a lower Φ and a higher Ef; such energy level alignment facilitates the transfer of photogenerated electrons. Furthermore, Wu et al. prepared ZIS rich with VZn via a solvothermal method and fabricated the Pt/VZn-ZIS 3D nanoflower photocatalyst (Figure 6j) [74]. During the HER process, the quantum efficiency of 1.0% Pt/ VZn-ZIS reached 13.12% at 350 nm and 6.59% at 400 nm (Figure 6k). Platinum atoms occupied the zinc vacancies in ZIS to form Pt-S electron channels, and the ΔGH* significantly decreased from −0.611 eV to −0.321 eV, which effectively reduced the energy barrier of the HER (Figure 6l) and effectively promoted the reduction in protons by photogenerated electrons. Chong et al., based on the two-dimensional layered structural characteristics of zinc indium sulfide (ZIS) and the specific coordination effect between ethylene glycol and Zn2+, incorporated ethylene glycol and zinc vacancies into the intra-layer and inter-layer of ZIS [75]. VZn, in synergy with the doped ions, form a strong built-in electric field, promoting the directional migration of photogenerated electron–hole pairs. In addition, VZn can serve as an H+ adsorption site, reducing the energy barrier of the HER and enhancing the absorption range of visible light.
Compared with Zn2+ ions, In3+ ions with more positive charges are easier to combine with small molecule chelators. Normally, it is easier for Zn2+ ions to form tetrahedral-possible zinc complexes with poor stabilities, while it is easier for In3+ ions to form more stable octahedral chelates of indium complex compounds. The addition of small molecule chelators could produce an In vacancy (VIn) in ZIS. Therefore, He et al. prepared a rich VIn- and poor VIn-ZIS via the hydrothermal process by adjusting the temperature from 393 to 473 K every 20 K using excess cysteine (C3H7NO2S) as the chelating agent [76]. The EPR signal intensities at the g-value of 2.005 indicated that the VIn-rich ZIS possessed a higher concentration of VIn than the VIn-poor ZIS (See Figure 7a–c).
Studies on dual vacancies in ZIS for the PHER (HER) have focused on their regulatory role in catalytic performances. As shown in Figure 7d–f, Zhao et al. successfully introduced dual Zn and In vacancies into ZIS flower like photocatalysts via cetyltrimethylammonium chloride (CTAC)-assisted hydrothermal method, where CTAC was used to induce structural defects [77]. A series of ZIS catalysts with different bimetallic defect concentrations were fabricated in the presence of different amounts of CTAC, which denoted them as ZIS-x. The increased intensity of the EPR signal in Figure 7f at a g-factor of 2.003 indicated increased vacancy concentrations in the order of ZIS < ZIS-1 < ZIS-5 < ZIS-10. As shown in Figure 7g, the Zn and In vacancies in ZIS-5 delocalize electrons and lead to charge accumulation around the defect sites from the charge density structures of ZIS and ZIS-5. As shown in Figure 7h–k, Zhong et al. investigated the effect of the concentrations of indium (In) and sulfur (S) vacancies on the photocatalytic activity of non-centrosymmetric zinc indium sulfide (ZIS) nanosheets in the HER, and a positive correlation was observed between the concentration of dual In and S vacancies and the PHER rate on ZIS nanosheets [78]. The synergistic effect of the dual vacancies increases the voltage gradient within the nanosheets and improves the charge separation efficiency after light absorption, thereby significantly enhancing the HER. Overall, these studies collectively demonstrate that different types of dual vacancies (Zn/In and In/S) in ZIS regulate the PHER performances through distinct yet effective mechanisms, and both can ultimately achieve the goal of promoting the HER.

2.3. Doping Engineering (Metal/Non-Metal Doping)

Doping engineering in ZIS, as a core strategy to tune material properties, achieves electronic structure optimization primarily through the strategic implantation of metal/non-metal elements, which effectively creates intermediate bandgap states to regulate electron transfer and energy band alignment [31,79,80]. Transition metals (such as Mn, Co, Ni, Zn, and Fe) can introduce intermediate energy levels that match the conduction bands/valence bands of ZIS, thereby adjusting the electron transition paths [81,82,83,84,85]; noble metals (such as Au, Ag, and Pt; often as nanoparticles formed on the defect sites) can not only form intermediate bandgap states but also synergistically improve carrier utilization efficiency through the surface plasmon resonance effect [86,87]. In addition, other common metals (like Cu and Na) can regulate the electronic transmission characteristics of ZIS by modifying the crystal lattice environment and inducing charge compensation effects, which also contributes to the optimization of the material’s overall electronic structure alongside the intermediate bandgap states [88]. Non-metal doping (e.g., N, S, and P) mainly modifies the band edge positions of ZIS, and when combined with the constructed intermediate bandgap states, it optimizes the range of light absorption and the kinetics of electron transmission [89,90]. Additionally, bimetallic co-doping—including combinations of transition metals (e.g., Co-Ni), noble metals (e.g., Au-Ag and Au-Pt), or transition metals with precious metals—leverages the electronic interaction between different metal ions to achieve precise regulation of intermediate bandgap states, further promoting the separation efficiency of photogenerated carriers [91,92]. In essence, all these doping modes revolve around optimizing the electronic structure of ZIS, laying a foundation for improving its photocatalytic performance in fields.
Noble metals are often used to enhance the performance of photocatalysts. He et al. prepared different structured Pt deposited ZIS photocatalysts, achieving an efficient HER. Under photothermal catalytic conditions (80 °C), the HER amount of 1% Pt/ZIS-300 reached 1218 μmol/(g catalyst) after 4 h of reaction [86]. However, their large-scale application is limited due to scarcity and high costs. Under this constraint, the design of transition metal-doped photocatalysts has become increasingly urgent. As shown in Figure 8a,b, Shen et al. investigated ZIS doped with different transition metals (Cr, Mn, Fe, and Co), and Mn-doped ZIS showed great improvement in the PHER, which may be attributed to the bandgap narrowing caused by Mn doping [93]. However, for Cr-, Fe-, and Co-doped ZIS, the inhibition of their photocatalytic activity should be ascribed to the dopant-related impurity energy levels; these levels can either cause the localization of charge carriers or act as non-radiative recombination centers for photoexcited electrons and holes. Summarily, He et al. achieved the goal of improving the PHER by Mn2+ doping and applying an external magnetic field [85]. Compared with ZIS, the H2 production rate of Mn0.15-ZIS reached as high as 32.75 mmol·g−1·h−1, which was 13.87 times that of ZIS. This demonstrates great potential in regulating spin-polarized electrons and provides an effective strategy for enhancing photocatalytic performances. As shown in Figure 8c–e, Xue et al. reported cobalt-doped ZIS hierarchical nanotubes (Co-ZIS HNTs) with bifunctions for the HER and benzaldehyde production [82]. Without any cocatalysts, the HER rate of Co-ZIS HNTs reached 2823.7 μmol·h−1·g−1, which was 1.93 times that of undoped ZIS microspheres (ZIS MSs), which may be according to the enhanced separation and transfer efficiency of photoinduced carriers in Co-ZIS HNTs.
Metal doping can regulate charge separation and construct gradient channels for hydrogen migration, thereby enhancing the performance of the PHER. As shown in Figure 8f–h, copper doping generates adaptive VS and establishes gradient hydrogen migration channels, reducing hydrogen diffusion resistance by 62% and collectively enhancing charge dynamics and HER efficiency [94]. The intensity of EPR signals at a 2.003 g-value in xCu–ZIS samples were related to the self-adapting vs. induced with atomic Cu introduction, which can tune charge separation and construct a gradient channel for H migration. Therefore, all samples exhibit photocatalytic H2 production under irradiation, with undoped ZIS showing a baseline rate of 0.6640 mmol·g−1·h−1. The H2 evolution rate increases with Cu dopant concentration, peaking at 9.8647 mmol·g−1·h−1 for 5%Cu-ZIS. Zeng et al. synthesized Sn-doped ZIS nanosheets via a one-step hydrothermal method to realize microstructural optimization, electronic property regulation, and active site modification of two-dimensional (2D) hexagonal-phase ZIS. The HER rate of Sn-ZIS under visible light reaches 62.18 μmol/h [96]. Also, for the modification of ZIS with noble metals and the introduction of defects present another effective regulatory approach, He et al. demonstrated that Ag-modified ZIS ultrathin nanosheets rich with VS could promote photocatalytic performances, which enriched the functional diversity of ZIS-based photocatalysts through structural and surface engineering [54]. Bimetallic co-doping to form photocatalysts with efficient charge separations and transfer performances is also crucial for a high-efficiency PHER. As shown in Figure 8i–l, Zhu et al. constructed a ZIS catalyst co-doped with Sn2+ and Sn4+ (Sn2+/Sn4+-ZIS) via a one-step solvothermal method [95]. Surface photovoltage (SPV) spectra show that all samples exhibit a strong response within the wavelength between 300 nm and 500 nm, and the Sn2+/Sn4+-ZIS sample has the strongest SPV intensity, indicating that the carrier separation efficiency is much higher than those of other catalysts; thus, the Sn2+/Sn4+-ZIS catalyst exhibited excellent HER catalytic activity (2.48 mmol·g−1·h−1). This strategy enhanced the metallic conductivity of ZIS and reduced the recombination of photogenerated electrons and holes. The co-doping of Sn2+ and Sn4+ ions increased the charge density and narrowed the bandgap for light harvesting. Bimetallic co-doping opens up a promising avenue for the development of high-efficiency photocatalysts for a solar-driven PHER. An et al. constructed a three-dimensional AgXAu1−X alloy/ZIS system (AgXAu1−X). In this system, AgXAu1−X alloy nanoparticles (NPs) are deposited on ZIS microspheres self-assembled from ZIS nanosheets, forming tight contact with ZIS. The HER capacity of Ag0.6Au0.4/ZIS was 7.1 times that of pure ZIS, and the apparent quantum yield (AQY) of Ag0.6Au0.4/ZIS was approximately 8.06% [91]. Considering the collective excitation of plasmonic nanoparticle assemblies, An et al. constructed a Au@Pt/ZIS photocatalyst through the assembly of core–shell structured Au@Pt nanoparticles on three-dimensional (3D) ZIS microspheres. This photocatalyst exhibited extraordinary catalytic performance in the water splitting for the HER under visible light (≥420 nm) [92]. Under visible light, the HER amount and HER rate of Au16@Pt/ZIS reached 41747 μmol·g−1 and 4174.7 μmol·g−1·h−1, respectively, which were approximately 10 times those of ZIS. Despite having achieved significant progress, doping engineering still faces some persistent challenges and requires further regulation.
Besides metal doping, non-metal doping can induce dual p-n charge characteristics within a single ZIS crystal structure, triggering charge transfer behavior, and promoting the PHER. As shown in Figure 9a, Chong et al. have demonstrated that the p-type characteristics induced by N doping lead to favorable charge redistribution in the ZIS framework [89]. N-doped ZIS (N-ZIS) showed a blue shift in the UV-vis spectra, accompanied by the dopant-induced color change (Figure 9b). The optimal 10N-ZIS remarkably boosts HER performance with the rate of 187.17 μmol∙g−1∙h−1, about 1575.71 μmol∙g−1 that of H2 after 6 h of visible-light irradiation (Figure 9c,d). To further investigate the changes in electronic properties and charge behavior induced by N doping, we computed the DOS and partial DOS (PDOS). As shown in Figure 9e, the valence band (VB) of pristine ZIS is composed of hybridized S 3p and Zn 3d orbitals, while the conduction band (CB) consists of S 3p and In 5s orbitals. The incorporation of N dopants into the ZIS lattice introduces a deep acceptor level above the valence band maximum. As a result, the Fermi level of N(4)(4’)ZIS shifts downward toward the VB, indicating a p-type conduction behavior. Moreover, N doping significantly enhances the electron density near the valence band maximum (VBM), suggesting a larger pool of ground-state electrons that can be photoexcited into the conduction band, thereby facilitating participation in the PHER. As shown in Figure 9f–i, Feng et al. reported that for phosphorus (P)-doped ZIS for the PHER, P-ZIS showed a good light absorption ability in the visible-light region, and the absorption edge of P-ZIS-1.0 was a little broader than that of the pristine ZIS, corresponding to its smaller bandgap values [97]. Thus, the P-ZIS exhibited stronger catalytic activity than the pristine ZIS, the HER rate of the optimized P-ZIS was 1566.6 μmol·g−1·h−1, which was 3.8 times that of the ZIS (411.1 μmol·g−1·h−1). As present in Figure 9j–m, Goswami et al. spectroscopically investigated the effects of O and N doped ZIS for the PHER [98]. OZIS and NZIS (ZIS doped with oxygen and nitrogen, respectively) exhibited reduced bandgap values. Furthermore, N and O doping resulted in an upward shift of the conduction band minimum (CBM) by approximately 0.06 eV and 0.21 eV for NZIS and OZIS, respectively, compared with pristine ZIS. The broad PL spectrum observed in the red region for OZIS suggests the existence of closely spaced defect states within the bandgap. Following O doping, the PL intensity undergoes an approximately 2-fold increase, which is likely attributed to the increased number of charge carriers in the system. In the doped ZIS system, the defect-state-mediated HER enabled the O-doped ZIS to result in a three-fold increase in the PHER rate. Also, Zhang et al. synthesized N, O co-doped ZnIn2S4 (NO-ZIS) and the N and O co-doping enhanced light absorption, increased specific surface area, and improved charge separation and transfer compared with unmodified ZIS [99]. As a result, the as-prepared NO-ZIS exhibited superior photocatalytic performance in both H2 evolution and Cr(VI) photoreduction relative to bare ZIS. Specifically, the H2 evolution rate of NO-ZIS reached 2254.0 μmol·g−1·h−1, which is 2.8 times that of pristine ZIS. In addition, the metal and non-metal co-doped (such as Mo/N, Ni/N, Co/P) ZIS system were also reported with enhanced PHER performances [100,101]. These findings collectively demonstrate the versatility of doping engineering in optimizing ZIS-based photocatalysts.

2.4. Performance Enhancements and Mechanistic Insights

The defect engineering strategies of ZIS effectively enhance photogenerated carrier separation and surface reaction kinetics through atomic-level modification. Based on modification mechanisms and experimental data, we construct a quantitative comparison in Table 1 as follows.

3. Conclusions and Outlooks

Although defect engineering strategies have significantly improved the PHER performance of ZIS as summarized in this review, several critical scientific and technological challenges remain.
(1) The real-time tracking of dynamic evolution at the atomic-scale—particularly the migration of vacancy defects during catalysis—remains unresolved. This limitation obstructs a deeper understanding and precise manipulation of the structure–performance relationship.
(2) For scalable applications, the high cost of noble-metal cocatalysts and the insufficient efficiency of non-precious alternatives present two major obstacles to practical implementation.
(3) The design of bifunctional active sites that can concurrently promote hydrogen evolution along with sacrificial agent oxidation, operate in tandem with the oxygen evolution reaction (OER) for overall water splitting, or facilitate synergistic organic pollutant degradation remains a significant challenge in system integration [110,111].
In the future, bridging the gap between laboratory research and industrial deployment will require integrated efforts across computational modeling, atomic-precision synthesis, and reactor engineering. With guidance from dynamic operando studies enabling precise defect construction, ZIS-based photocatalysts are poised to contribute substantially to the development of scalable solar-driven hydrogen production.

Author Contributions

Conceptualization: F.H., S.W. and Z.H.; writing—original draft preparation: F.H. and T.J.; writing—review and editing: S.W. and Z.H.; supervision: S.W. and Z.H.; funding acquisition: Z.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported under the framework of the National Natural Science Foundation of China (No. 22278245), Future Young Scholars Program of Shandong University (No. 61440089964189).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. He, Z.; Zhang, J.; Li, X.; Guan, S.; Dai, M.; Wang, S. 1D/2D heterostructured photocatalysts: From design and unique properties to their environmental applications. Small 2020, 16, 2005051. [Google Scholar] [CrossRef]
  2. Zhu, K.; Jie, O.-Y.; Zeng, Q.; Meng, S.; Teng, W.; Song, Y.; Tang, S.; Cui, Y. Fabrication of hierarchical ZnIn2S4OCNO nanosheets for photocatalytic hydrogen production and CO2 photoreduction. Chin. J. Catal. 2020, 41, 454–463. [Google Scholar] [CrossRef]
  3. Yu, H.; Dai, M.; Zhang, J.; Chen, W.; Jin, Q.; Wang, S.; He, Z. Interface engineering in 2D/2D heterogeneous photocatalysts. Small 2023, 19, 2205767. [Google Scholar] [CrossRef]
  4. Tian, S.; Feng, Y.; Zheng, Z.; He, Z. TiO2-Based photocatalytic coatings on glass substrates for environmental applications. Coatings 2023, 13, 1472. [Google Scholar] [CrossRef]
  5. Xiao, L.; Zhang, J.; Lu, T.; Zhou, G.-h.; Ren, Y.; Zheng, Z.; Yuan, X.-z.; Wang, S.-g.; He, Z. High-strength TiO2/TPU composite fiber based textiles for organic pollutant removal. npj Clean Water 2024, 7, 98. [Google Scholar] [CrossRef]
  6. Li, X.; Chang, J.; Zhang, S.; Xiao, L.; Wu, X.; He, Z. Microcystis@TiO2 nanoparticles for photocatalytic reduction reactions: Nitrogen fixation and hydrogen evolution. Catalysts 2021, 11, 1443. [Google Scholar] [CrossRef]
  7. Chang, Y.-C.; Chiao, Y.-C.; Hsu, P.-C. Rapid Microwave-Assisted synthesis of ZnIn2S4 nanosheets for highly efficient photocatalytic hydrogen production. Nanomaterials 2023, 13, 1957. [Google Scholar] [CrossRef] [PubMed]
  8. Yan, Y.; Meng, Q.; Tian, L.; Cai, Y.; Zhang, Y.; Chen, Y. Engineering of g-C3N4 for Photocatalytic Hydrogen Production: A Review. Int. J. Mol. Sci. 2024, 25, 8842. [Google Scholar] [CrossRef]
  9. Zhou, Z.; Jin, Z. Custom exposed crystal facets: Synergistic effect of optimum crystal facet anisotropy and Ohmic heterojunction boosting photocatalytic hydrogen evolution. Chin. J. Catal. 2025, 74, 294–307. [Google Scholar] [CrossRef]
  10. He, Z.; Kim, C.; Lin, L.H.; Jeon, T.H.; Lin, S.; Wang, X.C.; Choi, W. Formation of heterostructures via direct growth CN on h-BN porous nanosheets for metal-free photocatalysis. Nano Energy 2017, 42, 58–68. [Google Scholar] [CrossRef]
  11. He, Z.; Kim, C.; Jeon, T.H.; Choi, W. Hydrogenated heterojunction of boron nitride and titania enables the photocatalytic generation of H2 in the absence of noble metal catalysts. Appl. Catal. B Environ. 2018, 237, 772–782. [Google Scholar] [CrossRef]
  12. Hu, Q.; Niu, J.; Zhang, K.-Q.; Yao, M. One-dimensional CdS/SrTiO3/carbon fiber core-shell photocatalysts for enhanced photocatalytic hydrogen evolution. Coatings 2022, 12, 1235. [Google Scholar] [CrossRef]
  13. Zhang, W.; Yao, B.; Yang, H.; Li, X.; Qiu, L.; Li, S. Application of metals and their compounds/black phosphorus-based nanomaterials in the direction of photocatalytic hydrogen evolution. Coatings 2024, 14, 1141. [Google Scholar] [CrossRef]
  14. Jin, Z.; Li, Y.; Hao, X. Ni, Co-based selenide anchored g-C3N4 for boosting photocatalytic hydrogen evolution. Acta Phys.-Chim. Sin. 2021, 37, 1912033. [Google Scholar]
  15. Chen, L.-J.; Liu, T.-T.; Liu, S.-M.; Cai, S.; Zou, X.-X.; Jiang, J.-W.; Mei, Z.-Y.; Zhao, G.-F.; Yang, X.-F.; Guo, H. S vacant CuIn5S8 confined in a few-layer MoSe2 with interlayer-expanded hollow heterostructures boost photocatalytic CO2 reduction. Rare Met. 2022, 41, 144–154. [Google Scholar] [CrossRef]
  16. Sun, T.; Li, C.; Bao, Y.; Fan, J.; Liu, E. S-scheme MnCo2S4/g-C3N4 heterojunction photocatalyst for H2 production. Acta Phys.-Chim. Sin. 2023, 39, 2212009. [Google Scholar] [CrossRef]
  17. Wang, R.-Z.; Lin, Z.; Wang, Y.-Q.; Zhang, K.-N.; Zhang, G.-H.; Zhang, J.; Mao, S.S.; Shen, S.-H. A direct polymeric carbon nitride/tungsten oxide Z-scheme heterostructure for efficient photocatalytic hydrogen generation via reforming of plastics into value-added chemicals. Rare Met. 2024, 43, 3771–3783. [Google Scholar] [CrossRef]
  18. Xiao, L.H.; Li, X.; Zhang, J.; He, Z.L. MgB4 MXene-like nanosheets for photocatalytic hydrogen evolution. ACS Appl. Nano Mater. 2021, 4, 12779–12787. [Google Scholar] [CrossRef]
  19. Li, M.; Wang, J.-Z.; Jin, Z.-L. Graphdiyne (CnH2n-2) as an “electron transfer bridge” boosting photocatalytic hydrogen evolution over Zn0.5Co0.5S/MoS2 S-scheme heterojunction. Rare Met. 2024, 43, 1999–2014. [Google Scholar] [CrossRef]
  20. Liu, X.-Y.; Cao, Q.; Li, G.-X.; Liu, H.; Zeng, L.-L.; Zhao, L.-L.; Chang, B.; Wang, X.-W.; Liu, H.; Zhou, W.-J. Orientation controlled photogenerated carriers on self-supporting CdS/Ni3S2 paper toward photocatalytic hydrogen evolution and biomass upgrading. Rare Met. 2024, 43, 2015–2025. [Google Scholar] [CrossRef]
  21. Zhang, Z.; Zhang, J.; Wang, H.; Liu, M.; Xu, Y.; Liu, K.; Zhang, B.; Shi, K.; Zhang, J.; Ma, G. Facet-oriented surface modification for enhancing photocatalytic hydrogen production on Sm2Ti2O5S2 nanosheets. Chin. J. Catal. 2025, 74, 341–351. [Google Scholar] [CrossRef]
  22. Dai, M.; He, Z.; Zhang, P.; Li, X.; Wang, S. ZnWO4-ZnIn2S4 S-scheme heterojunction for enhanced photocatalytic H2 evolution. J. Mater. Sci. Technol. 2022, 122, 231–242. [Google Scholar] [CrossRef]
  23. Cai, Y.; Luo, F.; Guo, Y.; Guo, F.; Shi, W.; Yang, S. Near-infrared light driven ZnIn2S4-based photocatalysts for environmental and energy applications: Progress and perspectives. Molecules 2023, 28, 2142. [Google Scholar] [CrossRef] [PubMed]
  24. Li, W.; Li, J.-J.; Liu, Z.-F.; Ma, H.-Y.; Fang, P.-F.; Xiong, R.; Wei, J.-H. Fast charge transfer kinetics in Sv-ZnIn2S4/Sb2S3 S-scheme heterojunction photocatalyst for enhanced photocatalytic hydrogen evolution. Rare Met. 2024, 43, 533–542. [Google Scholar] [CrossRef]
  25. Yang, W.-N.; Yang, J.; Yang, H.; Sun, L.; Li, H.-X.; Li, D.-C.; Dou, J.-M.; Li, X.-G.; Cao, G.-D. In-situ construction of tubular core-shell noble-metal-free CMT@TiO2/ZnIn2S4 S-scheme heterojunction for superior photothermal-photocatalytic hydrogen evolution. Rare Met. 2025, 44, 2474–2488. [Google Scholar] [CrossRef]
  26. Dai, M.; He, Z.; Cao, W.; Zhang, J.; Chen, W.; Jin, Q.; Que, W.; Wang, S. Rational construction of S-scheme BN/MXene/ZnIn2S4 heterojunction with interface engineering for efficient photocatalytic hydrogen production and chlorophenols degradation. Sep. Purif. Technol. 2023, 309, 123004. [Google Scholar] [CrossRef]
  27. Zheng, X.L.; Song, Y.M.; Liu, Y.H.; Yang, Y.Q.; Wu, D.X.; Yang, Y.J.; Feng, S.Y.; Li, J.; Liu, W.F.; Shen, Y.J.; et al. ZnIn2S4-based photocatalysts for photocatalytic hydrogen evolution via water splitting. Coord. Chem. Rev. 2023, 475, 214898. [Google Scholar] [CrossRef]
  28. Dai, M.; Yu, H.; Chen, W.; Qu, K.-A.; Zhai, D.; Liu, C.; Zhao, S.; Wang, S.; He, Z. Boosting photocatalytic activity of CdLa2S4/ZnIn2S4 S-scheme heterojunctions with spatial separation of photoexcited carries. Chem. Eng. J. 2023, 470, 144240. [Google Scholar] [CrossRef]
  29. Zheng, X.; Song, Y.; Wang, C.; Gao, Q.; Shao, Z.; Lin, J.; Zhai, J.; Li, J.; Shi, X.; Wu, D.; et al. Properties, applications, and challenges of copper- and zinc-based multinary metal sulfide photocatalysts for photocatalytic hydrogen evolution. Chin. J. Catal. 2025, 74, 22–70. [Google Scholar] [CrossRef]
  30. Du, C.; Zhang, Q.; Lin, Z.Y.; Yan, B.; Xia, C.X.; Yang, G.W. Half-unit-cell ZnIn2S4 monolayer with sulfur vacancies for photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2019, 248, 193–201. [Google Scholar] [CrossRef]
  31. Yadav, G.; Ahmaruzzaman, M. ZnIn2S4 and ZnIn2S4 based advanced hybrid materials: Structure, morphology and applications in environment and energy. Inorg. Chem. Commun. 2022, 138, 109288. [Google Scholar] [CrossRef]
  32. Zhong, Y.Q.; Li, M.Y.; Luan, X.; Gao, F.F.; Wu, H.X.; Zi, J.Z.; Lian, Z.C. Ultrathin ZnIn2S4/ZnSe heteronanosheets with modulated S-scheme enable high efficiency of visible-light-responsive photocatalytic hydrogen evolution. Appl. Catal. B Environ. Energy 2023, 335, 122859. [Google Scholar] [CrossRef]
  33. Yao, L.; Zeng, S.C.; Yang, S.X.; Zhang, H.H.; Ma, Y.; Zhou, G.Y.; Fang, J.Z. Zinc indium sulfide materials for photocatalytic hydrogen production via water splitting: A short review. Catalysts 2025, 15, 271. [Google Scholar] [CrossRef]
  34. Zhao, L.; Chang, T.; Hu, Z.; Fang, K.; Zhang, X.; Xiao, W.; Jiang, F.; Wang, L.; Liu, D.; Zhang, Y. Rapid carrier extraction and d-band center regulation of Pd-ZnIn2S4 for efficient photocatalytic water splitting. J. Alloys Compd. 2025, 1027, 180652. [Google Scholar] [CrossRef]
  35. Wu, X.; Qian, Y.; Lv, G.; Long, L.; Zhou, Y.; Wang, D. Realizing efficient photoelectrochemical performance for well-designed CdS@ZnIn2S4 heterostructure photoanode with directional interfacial charge transfer dynamics. Coatings 2022, 12, 1210. [Google Scholar] [CrossRef]
  36. Wang, B.; Ding, Y.; Deng, Z.; Li, Z. Rational design of ternary NiS/CQDs/ZnIn2S4 nanocomposites as efficient noble-metal-free photocatalyst for hydrogen evolution under visible light. Chin. J. Catal. 2019, 40, 335–342. [Google Scholar] [CrossRef]
  37. Li, X.-J.; Qi, M.-Y.; Li, J.-Y.; Tan, C.L.; Tang, Z.-R.; Xu, Y.-J. Visible light-driven dehydrocoupling of thiols to disulfides and H2 evolution over PdS-decorated ZnIn2S4 composites. Chin. J. Catal. 2023, 51, 55–65. [Google Scholar] [CrossRef]
  38. Rahman, M.H.; Yang, J.; Sun, Y.; Mannodi-Kanakkithodi, A. Defect engineering in ZnIn2X4 (X = S, Se, Te) semiconductors for improved photocatalysis. Surf. Interf. 2023, 39, 102960. [Google Scholar] [CrossRef]
  39. Li, J.; Wu, C.; Li, J.; Dong, B.; Zhao, L.; Wang, S. 1D/2D TiO2/ZnIn2S4 S-scheme heterojunction photocatalyst for efficient hydrogen evolution. Chin. J. Catal. 2022, 43, 339–349. [Google Scholar] [CrossRef]
  40. Zhao, M.; Liu, S.; Chen, D.; Zhang, S.; Carabineiro, S.A.C.; Lv, K. A novel S-scheme 3D ZnIn2S4/WO3 heterostructure for improved hydrogen production under visible light irradiation. Chin. J. Catal. 2022, 43, 2615–2624. [Google Scholar] [CrossRef]
  41. Li, W.; Yang, Z.; Li, Y.; Zhang, P.; Li, H. Highly efficient photocathodic protection performance of ZIS@CNNs composites under visible light. Coatings 2023, 13, 1479. [Google Scholar] [CrossRef]
  42. Li, W.-Y.; Liu, J.; Li, Z.-R.; Shen, T.-T.; Chen, J.; Hu, Z.-Y.; Yu, W.-B.; Li, Y.; Su, B.-L. Probing the deactivation and regeneration of ZnIn2S4 in photocatalytic degradation of tetracycline. J. Colloid Interface Sci. 2025, 694, 137700. [Google Scholar] [CrossRef]
  43. Zhang, K.; Dan, M.; Yang, J.; Wu, F.; Wang, L.; Tang, H.; Liu, Z. Surface energy mediated sulfur vacancy of ZnIn2S4 atomic layers for photocatalytic H2O2 production. Adv. Funct. Mater. 2023, 33, 2302964. [Google Scholar] [CrossRef]
  44. Zou, P.; Wu, Z.; Ma, S.; Cao, G.; Jiang, X.; Wang, H. Crystal plane regulation and heterostructure construction of ZnIn2S4/h-BN for boosting photocatalytic hydrogen evolution. Appl. Surf. Sci. 2025, 697, 163017. [Google Scholar] [CrossRef]
  45. Tan, M.; Huang, C.; Yu, C.; Li, C.; Yin, R.; Liu, C.; Dong, W.; Meng, H.; Su, Y.; Qiao, L.; et al. Unexpected high-performance photocatalytic hydrogen evolution in Co@NCNT@ZnIn2S4 triggered by directional charge separation and transfer. Small 2022, 18, 2205266. [Google Scholar] [CrossRef]
  46. Fan, H.; Wu, Z.; Liu, K.; Liu, W. Fabrication of 3D CuS@ZnIn2S4 hierarchical nanocages with 2D/2D nanosheet subunits p-n heterojunctions for improved photocatalytic hydrogen evolution. Chem. Eng. J. 2022, 433, 134474. [Google Scholar] [CrossRef]
  47. Li, X.; Huang, Y.; Ho, W.; Han, S.; Wang, P.; Lee, S.; Zhang, Z. Modulation of sulfur vacancies at ZnIn2S4-δ/g-C3N4 heterojunction interface for successive C-H secession in photocatalytic gaseous formaldehyde complete oxidation. Appl. Catal. B Environ. 2023, 338, 123048. [Google Scholar] [CrossRef]
  48. Sun, R.; Liu, Y.; Yang, J.; Wuren, T.; Duan, H.; Tan, Z.; Yu, S. Mo-W18O49/ZnIn2S4 composites synthesized by metal doping for photocatalytic hydrogen evolution. Molecules 2025, 30, 1563. [Google Scholar] [CrossRef]
  49. Shi, W.L.; Chen, Z.Z.; Lu, J.L.; Sun, X.H.; Wang, Z.Y.; Yan, Y.J.; Guo, F.; Chen, L.Z.; Wang, G.Z. Construction of ZrC@ZnIn2S4 core-shell heterostructures for boosted near-infrared-light driven photothermal-assisted photocatalytic H2 evolution. Chem. Eng. J. 2023, 474, 145690. [Google Scholar] [CrossRef]
  50. Chen, X.; Liang, B.; Yan, L. Improvement in the photocatalytic hydrogen production of flower-shaped ZnIn2S4 by surface modification with amino silane. Catalysts 2024, 14, 607. [Google Scholar] [CrossRef]
  51. Wang, S.; Cheng, Y.; Huang, W.; Dou, M.; Shao, H.; Yao, M.; Ding, K.; Ye, T.; Zhou, R.; Li, S.; et al. The Zn vacancy-mediated de-accumulation based process for hydrogen production performance promotion of 1D Zn─Cd─S nanorods. Small 2023, 20, 2306447. [Google Scholar] [CrossRef] [PubMed]
  52. Yadav, N.; Pant, K.K.; Tripathi, K.; Yadav, G.; Ahmaruzzaman, M. Harnessing metal sulfides for efficient hydrogen production employing photocatalytic water splitting: Current status and future direction. Int. J. Hydrogen Energy 2025, 101, 1221–1253. [Google Scholar] [CrossRef]
  53. Zhang, N.; Xing, Z.; Li, Z.; Zhou, W. Sulfur vacancy engineering of metal sulfide photocatalysts for solar energy conversion. Chem. Catal. 2023, 3, 100375. [Google Scholar] [CrossRef]
  54. He, Z.; Li, L.; Pang, Y.; Cui, M.; Lin, Y.; Xie, T. Ag-modified sulfur vacancy-rich ZnIn2S4 ultrathin Nanosheets promote oxygen activation for efficient photocatalytic hydrogen peroxide synthesis. J. Colloid. Interface Sci. 2025, 700, 138353. [Google Scholar] [CrossRef]
  55. 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]
  56. Jing, H.T.; Ren, J.; Yue, J.Y.; Liu, S.Y.; Liang, Q.F.; Wu, R.; Wang, Y.W.; Fang, Z.B.; Li, H.L.; Wei, S.H. 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]
  57. Wang, Y.; Zhang, H.H.; Wen, Z.; Sun, C.N.; Wang, G.Y.; Wang, M.-S.; Yang, C.C.; Jiang, Q. ZnIn2S4 with a hybrid reaction mechanism and sulfur vacancies for sustainable sodium storage. Carbon. Energy 2025, 7, e654. [Google Scholar] [CrossRef]
  58. Huang, M.; Zhang, H.; Xu, M.; Chen, W.-J.; Pan, X.; Liang, S. Controllable engineering of ZnIn2S4 with sulfur vacancy as an efficient piezocatalyst toward H2 production. J. Alloys Compd. 2025, 1016, 179005. [Google Scholar] [CrossRef]
  59. He, J.; Yang, Z.; Wang, Z.; Gu, L.; Qiu, J.; Ran, J. Sulfur-vacancy-enriched ZnIn2S4 mediates efficient charge transfer for hydrogen evolution from lignocellulose photoreforming. Chem. Eng. J. 2025, 503, 158433. [Google Scholar] [CrossRef]
  60. Tang, J.-P.; Chen, Y.; Wang, Z.-Y.; Hu, Y.-H.; Wang, J.-H.; Bao, L.; Zhao, Z.-Y.; Yuan, Y.-J. Sustainable H2 production from lignocellulosic biomass over MoS2 modified sulfur vacancy enriched ZnIn2S4 photocatalyst. ACS Catal. 2025, 15, 265–274. [Google Scholar] [CrossRef]
  61. Xu, M.H.; Ruan, X.W.; Meng, D.P.; Fang, G.Z.; Jiao, D.X.; Zhao, S.L.; Liu, Z.Y.; Jiang, Z.F.; Ba, K.K.; Xie, T.F.; et al. Modulation of sulfur vacancies in ZnIn2S4/MXene schottky heterojunction photocatalyst promotes hydrogen evolution. Adv. Funct. Mater. 2024, 34, 2402330. [Google Scholar] [CrossRef]
  62. Ren, S.; Wang, S.; Chen, H.; Li, M.; Li, Z.; Liao, L.; Liang, Z.; Zhou, W. S vacancies engineering ZnIn2S4 nanosheets for boosted photocatalytic hydrogen evolution. Int. J. Hydrogen Energy 2024, 51, 1128–1135. [Google Scholar] [CrossRef]
  63. Wang, X.H.; Wang, X.H.; Huang, J.F.; Li, S.X.; Meng, A.; Li, Z.J. Interfacial chemical bond and internal electric field modulated Z-scheme SV-ZnIn2S4/MoSe2 photocatalyst for efficient hydrogen evolution. Nat. Commun. 2021, 12, 4112. [Google Scholar] [CrossRef]
  64. Li, G.Q.; Liang, H.O.; Fan, X.Y.; Lv, X.L.; Sun, X.W.; Wang, H.G.; Bai, J. Modulating and optimizing 2D/2D Fe-Ni2P/ZnIn2S4 with S vacancy through surface engineering for efficient photocatalytic H2 evolution. J. Mater. Chem. A 2023, 11, 14809–14818. [Google Scholar] [CrossRef]
  65. Liu, S.; Man, Z.; Fang, F.; Li, P.; Chang, K. Controllable engineering of surface defects on ZnIn2S4 nanosheets: Implications to efficient hydrogen evolution activity under visible light. ACS Appl. Nano Mater. 2023, 6, 14876–14884. [Google Scholar] [CrossRef]
  66. Jiang, R.Q.; Cai, X.Y.; Gu, X.Q.; Yang, D.; Zhang, J.Y.; Zhao, Y.L.; Mao, L. Photocatalytic H2 evolution over sulfur vacancy-rich ZnIn2S4 hierarchical microspheres under visible light. J. Mater. Sci. 2021, 56, 19439–19451. [Google Scholar] [CrossRef]
  67. Chen, J.; Li, K.; Cai, X.; Zhao, Y.; Gu, X.; Mao, L. Sulfur vacancy-rich ZnIn2S4 nanosheet arrays for visible-light-driven water splitting. Mater. Sci. Semicond. Process. 2022, 143, 106547. [Google Scholar] [CrossRef]
  68. Hsia, H.-H.; Maggay, I.V.B.; Muruganantham, R.; Liu, W.-R. ZnIn2S4: A promising anode material with high electrochemical performance for sodium-ion batteries. Ceram. Int. 2021, 47, 28634–28641. [Google Scholar] [CrossRef]
  69. Roda, D.; Trzciński, K.; Sawczak, M.; Ilnicka, A.; Łapiński, M.; Szkoda, M.; Nowak, A.P. ZnIn2S4 thin films grown by pulsed laser deposition: Effect of electrolyte and illumination conditions on photoanode performance. Appl. Surf. Sci. 2025, 708, 163694. [Google Scholar] [CrossRef]
  70. He, Y.; Rao, H.; Song, K.; Li, J.; Yu, Y.; Lou, Y.; Li, C.; Han, Y.; Shi, Z.; Feng, S. 3D Hierarchical ZnIn2S4 Nanosheets with Rich Zn Vacancies Boosting Photocatalytic CO2 Reduction. Adv. Funct. Mater. 2019, 29, 1905153. [Google Scholar] [CrossRef]
  71. Long, L.; Lv, G.; Pan, F.; Li, Z.; Zhu, H.; Wang, D. Photocarrier dynamics regulation by constructing a 2D ZnIn2S4 nanosheet array with highly exposed Zn vacancies for efficient solar-driven CO2 conversion. J. Phys. Chem. C 2023, 127, 24077–24087. [Google Scholar] [CrossRef]
  72. Xu, B.; Li, H.; Chong, B.; Lin, B.; Yan, X.; Yang, G. Zn vacancy-tailoring mediated ZnIn2S4 nanosheets with accelerated orderly charge flow for boosting photocatalytic hydrogen evolution. Chem. Eng. Sci. 2023, 270, 118533. [Google Scholar] [CrossRef]
  73. Li, G.; Wang, X.; Li, H.; Jiang, Q.; Tang, S.; Pang, H.; Ma, H. Synergizing Zinc vacancy engineering and electron-rich polyoxometalates in ZnIn2S4 for solar-light-driven hydrogen evolution with boosted charge dynamics. J. Alloys Compd. 2025, 1036, 182085. [Google Scholar] [CrossRef]
  74. Wu, W.D.; Luo, Z.C.; Liu, B.W.; Qiu, X.Q.; Lin, J.X.; Sun, S.R.; Wang, X.F.; Lin, X.L.; Qin, Y.L. Zinc vacancy promotes photo-reforming Ligninmodel to H2 evolution and value-added chemicals production. Small Methods 2023, 7, 2300462. [Google Scholar] [CrossRef] [PubMed]
  75. Ouyang, C.; Quan, X.; Chen, Z.A.; Wang, K.; Gu, X.; Hong, Z.; Zhi, M. Intercalation- and vacancy-enhanced internal electric fields in ZnIn2S4 for highly efficient photocatalytic H2O2 production. J. Phys. Chem. C 2023, 127, 20683–20699. [Google Scholar] [CrossRef]
  76. He, Y.; Chen, C.; Liu, Y.; Yang, Y.; Li, C.; Shi, Z.; Han, Y.; Feng, S. Quantitative evaluation of carrier dynamics in full-spectrum responsive metallic ZnIn2S4 with indium vacancies for boosting photocatalytic CO2 reduction. Nano Lett. 2022, 22, 4970–4978. [Google Scholar] [CrossRef] [PubMed]
  77. Zhao, J.; Wang, Y.; Liu, H.; Zhang, R.; Jia, W.; Zhang, J.; Sun, Y.; Peng, L. Understanding the role of dual zinc and indium vacancies in ZnIn2S4 for the visible-light-driven photocatalytic air-oxidation of 5-hydroxymethylfurfural. ACS Catal. 2025, 15, 3464–3474. [Google Scholar] [CrossRef]
  78. Zhong, W.J.; Hung, M.Y.; Kuo, Y.T.; Tian, H.K.; Tsai, C.N.; Wu, C.J.; Lin, Y.D.; Yu, H.C.; Lin, Y.G.; Wu, J.J. Dual-vacancy-engineered ZnIn2S4 nanosheets for harnessing low-frequency vibration induced piezoelectric polarization coupled with static dipole field to enhance photocatalytic H2 evolution. Adv. Mater. 2024, 36, 2403228. [Google Scholar] [CrossRef]
  79. Xiao, L.; Zhang, S.; Cui, H.; Chang, J.; Feng, Y.; Wang, S.; He, Z. Promoted photocatalytic performances over Ti3+-B co-doped TiO2/BN with high carrier transfer and absorption capabilities driven by SWCNT addition. Mater. Sci. Semicond. Process. 2024, 177, 108364. [Google Scholar] [CrossRef]
  80. Su, Y.; Li, C.; Jiang, X.; Bai, Y. First-principles study of non-metallic doping and noble metal loading on ZnIn2S4 semiconductor photocatalyst. Inorg. Chem. Commun. 2025, 174, 114124. [Google Scholar] [CrossRef]
  81. Li, J.-Y.; Qi, M.-Y.; Xu, Y.-J. Efficient splitting of alcohols into hydrogen and C–C coupled products over ultrathin Ni-doped ZnIn2S4 nanosheet photocatalyst. Chin. J. Catal. 2022, 43, 1084–1091. [Google Scholar] [CrossRef]
  82. Xue, J.R.; Liu, H.H.; Zeng, S.Y.; Feng, Y.F.; Zhang, Y.Y.; Zhu, Y.; Cheng, M.Y.; Zhang, H.K.; Shi, L.; Zhang, G.Q. Bifunctional cobalt-doped ZnIn2S4 hierarchical nanotubes endow noble-metal cocatalyst-free photocatalytic H2 production coupled with benzyl alcohol oxidation. Sol. RRL 2022, 6, 2101042. [Google Scholar] [CrossRef]
  83. Zhang, S.; Dai, M.; Guo, J.; Wang, G.; Wang, S.; He, Z. Stable Ti3+ in B-TiO2/BN based hybrids for efficient photocatalytic reduction. Chem. Eng. J. Adv. 2022, 11, 100333. [Google Scholar] [CrossRef]
  84. Li, M.; Chen, L.; Xiao, Q.; Yang, Y.; Chen, H.; Qiu, X. Self-doped zinc-mediated ZnIn2S4 microflowers towards optimized visible-light photocatalytic hydrogen production. J. Environ. Chem. Eng. 2024, 12, 112844. [Google Scholar] [CrossRef]
  85. He, Z.; Xia, Y.; He, G.; Wang, F.; Su, J.; Fareed, H.; Fu, X.; Chen, G.; Yang, H.; Zhou, W. Insight into synergy of Mn active sites and spin polarization electrons in Mn-incorporated ZnIn2S4 for boosting photocatalytic hydrogen evolution coupled with benzyl alcohol oxidation. Chem. Eng. J. 2025, 506, 159957. [Google Scholar] [CrossRef]
  86. He, J.; Yang, Z.; Wang, Z.; Qiu, J.; Ran, J. Photothermal catalysis enhances cellulose oxidation kinetics to promote the hydrogen evolution in alkaline solution over Pt/ZnIn2S4. Energy Convers. Manag. 2024, 314, 118634. [Google Scholar] [CrossRef]
  87. Yang, M.; Zhan, X.-Q.; Ou, D.-L.; Wang, L.; Zhao, L.-L.; Yang, H.-L.; Liao, Z.-Y.; Yang, W.-Y.; Ma, G.-Z.; Hou, H.-L. Efficient visible-light-driven hydrogen production with Ag-doped flower-like ZnIn2S4 microspheres. Rare Met. 2025, 44, 1024–1041. [Google Scholar] [CrossRef]
  88. Yuan, S.; Liu, G.; Zhang, Q.; Liu, T.; Yang, J.; Guan, Z. Synergistic effect of Na doping and CoSe2 cocatalyst for enhanced photocatalytic hydrogen evolution performance of ZnIn2S4. J. Colloid Interface Sci. 2024, 676, 272–282. [Google Scholar] [CrossRef]
  89. Chong, W.-K.; Ng, B.-J.; Kong, X.Y.; Tan, L.-L.; Putri, L.K.; Chai, S.-P. Non-metal doping induced dual p-n charge properties in a single ZnIn2S4 crystal structure provoking charge transfer behaviors and boosting photocatalytic hydrogen generation. Appl. Catal. B Environ. 2023, 325, 122372. [Google Scholar] [CrossRef]
  90. Farhan, S.; Hassan Raza, A.; Yang, S.; Yu, Z.; Wu, Y. Boosted photocatalytic hydrogen evolution of S-scheme N-doped CeO2-δ@ZnIn2S4 heterostructure photocatalyst. J. Colloid Interface Sci. 2024, 669, 430–443. [Google Scholar] [CrossRef]
  91. An, H.; Li, M.; Liu, R.; Gao, Z.; Yin, Z. Design of AgXAu1−X alloy/ZnIn2S4 system with tunable spectral response and Schottky barrier height for visible-light-driven hydrogen evolution. Chem. Eng. J. 2020, 382, 122953. [Google Scholar] [CrossRef]
  92. An, H.; Lv, Z.; Zhang, K.; Deng, C.; Wang, H.; Xu, Z.; Wang, M.; Yin, Z. Plasmonic coupling enhancement of core-shell Au@Pt assemblies on ZnIn2S4 nanosheets towards photocatalytic H2 production. Appl. Surf. Sci. 2021, 536, 147934. [Google Scholar] [CrossRef]
  93. Shen, S.; Chen, J.; Wang, X.; Zhao, L.; Guo, L. Microwave-assisted hydrothermal synthesis of transition-metal doped ZnIn2S4 and its photocatalytic activity for hydrogen evolution under visible light. J. Power Sources 2011, 196, 10112–10119. [Google Scholar] [CrossRef]
  94. Zhang, S.; Zhang, Z.; Si, Y.; Li, B.; Deng, F.; Yang, L.; Liu, X.; Dai, W.; Luo, S. Gradient hydrogen migration modulated with self-adapting S vacancy in copper-doped ZnIn2S4 nanosheet for photocatalytic hydrogen evolution. ACS Nano 2021, 15, 15238–15248. [Google Scholar] [CrossRef] [PubMed]
  95. Zhu, H.; Gu, S.; Zhang, C.; Gao, K.; Xia, J.; Cheng, C.; Wang, X. Regulating electron-transfer over ZnIn2S4 by Sn(II)/Sn(IV) co-doping for efficient photocatalytic hydrogen production. Catal. Lett. 2022, 153, 2245–2259. [Google Scholar] [CrossRef]
  96. Zeng, Z.; Mao, L.; Zhang, R.; Liu, Y.; Ling, Y.; Cai, X.; Zhang, J. A triple-benefit strategy of microstructure, electronic property and active site modulation on ZnIn2S4 photocatalyst by Sn atom doping. Chem. Eng. J. 2024, 496, 153769. [Google Scholar] [CrossRef]
  97. Feng, X.; Chen, H.; Yin, H.; Yuan, C.; Lv, H.; Fei, Q.; Zhang, Y.; Zhao, Q.; Zheng, M.; Zhang, Y. Facile synthesis of P-doped ZnIn2S4 with enhanced visible-light-driven photocatalytic hydrogen production. Molecules 2023, 28, 4520. [Google Scholar] [CrossRef]
  98. Goswami, T.; Yadav, D.K.; Bhatt, H.; Kaur, G.; Shukla, A.; Babu, K.J.; Ghosh, H.N. Defect-mediated slow carrier recombination and broad photoluminescence in non-metal-doped ZnIn2S4 nanosheets for enhanced photocatalytic activity. J. Phys. Chem. Lett. 2021, 12, 5000–5008. [Google Scholar] [CrossRef] [PubMed]
  99. Zhang, M.; Du, M.; Ren, Y.; Zhu, Y.; Dong, Y.; Liao, J.; Chen, X.; Zheng, Q.; Yin, H. One-step synthesis N, O co-doped ZnIn2S4 for efficient photocatalytic H2 production and Cr(VI) reduction. J. Environ. Chem. Eng. 2024, 12, 113261. [Google Scholar] [CrossRef]
  100. Liu, Q.; Wang, M.; He, Y.; Wang, X.; Su, W. Photochemical route for synthesizing Co–P alloy decorated ZnIn2S4 with enhanced photocatalytic H2 production activity under visible light irradiation. Nanoscale 2018, 10, 19100–19106. [Google Scholar] [CrossRef]
  101. Wu, M.; Sheng, Q.; Yang, L.; Lv, W.; Zhen, W.; Liu, H.; Yang, Y.; Yao, T.; Zhang, J.; Liu, L. Preparation of N,Ni co-doped ZnIn2S4 for simultaneously trapping electron and hole: Insight on photocatalytic H2 evolution coupled with Benzyl Alcohols oxidative dehydrogenation. Small 2025, 21, 2500443. [Google Scholar] [CrossRef]
  102. Yi, L.; Nkinahamira, F.; Jiang, H.; Huang, J.; Ma, Y.; Zhu, R. Mo2N decorated hierarchical ZnIn2S4 with sulfur-rich vacancy photocatalyst for efficient photocatalytic hydrogen production. J. Environ. Chem. Eng. 2023, 11, 110240. [Google Scholar] [CrossRef]
  103. Han, Q.; Wang, B.; Gao, J.; Qu, L. Graphitic carbon nitride/nitrogen-rich carbon nanofibers: Highly efficient photocatalytic hydrogen evolution without cocatalysts. Angew. Chem. Int. Ed. 2016, 55, 10849–10853. [Google Scholar] [CrossRef] [PubMed]
  104. Ran, Q.; Yu, Z.; Jiang, R.; Hou, Y.; Huang, J.; Zhu, H.; Yang, F.; Li, M.; Li, F.; Sun, Q. A novel, noble-metal-free core-shell structure Ni–P@C cocatalyst modified sulfur vacancy-rich ZnIn2S4 2D ultrathin sheets for visible light-driven photocatalytic hydrogen evolution. J. Alloys Compd. 2021, 855, 157333. [Google Scholar] [CrossRef]
  105. Liu, X.; Xu, J.; Jiang, Y.; Du, Y.; Zhang, J.; Lin, K. In-situ construction of CdS@ZIS Z-scheme heterojunction with core-shell structure: Defect engineering, enhance photocatalytic hydrogen evolution and inhibit photo-corrosion. Int. J. Hydrogen Energy 2022, 47, 35241–35253. [Google Scholar] [CrossRef]
  106. Wang, S.D.; Huang, L.Y.; Xue, L.J.; Kang, Q.; Wen, L.L.; Lv, K.L. Sulfur-vacancy-modified ZnIn2S4/TpPa-1 S-scheme heterojunction with enhanced internal electric field for boosted photocatalytic hydrogen production. Appl. Catal. B Environ. Energy 2024, 358, 124366. [Google Scholar] [CrossRef]
  107. Shi, X.; Dai, C.; Wang, X.; Hu, J.; Zhang, J.; Zheng, L.; Mao, L.; Zheng, H.; Zhu, M. Protruding Pt single-sites on hexagonal ZnIn2S4 to accelerate photocatalytic hydrogen evolution. Nat. Commun. 2022, 13, 1287. [Google Scholar] [CrossRef]
  108. Liu, W.; Chen, J.; Pan, X.; Wang, T.; Li, Y. Ultrathin nickel-doped ZnIn2S4 nanosheets with sulfur vacancies for efficient photocatalytic hydrogen evolution. ChemCatChem 2021, 13, 5148–5155. [Google Scholar] [CrossRef]
  109. Zeng, D.; Shen, T.; Hu, Y.; Zhang, Z.; Liu, Z.; Xu, N.; Song, J.; Guan, R.; Zhou, C. Nitrogen-doped ZnIn2S4 and TPA multi-dimensional synergistically enhance photocatalytic hydrogen production. Fuel 2025, 380, 133151. [Google Scholar] [CrossRef]
  110. Tien, T.-M.; Chen, E.L. S-scheme system of MoS2/Co3O4 nanocomposites for enhanced photocatalytic hydrogen evolution and methyl violet dye removal under visible light irradiation. Coatings 2023, 13, 80. [Google Scholar] [CrossRef]
  111. Cheng, T.; Zhu, J.; Chen, C.; Hu, Y.; Wu, L.; Zhang, M.; Cui, L.; Dai, Y.; Zhang, X.; Tian, Y.; et al. Construction of advanced S-scheme heterojunction interface composites of bimetallic phosphate MnMgPO4 with C3N4 surface with remarkable performance in photocatalytic hydrogen production and pollutant degradation. Coatings 2025, 15, 103. [Google Scholar] [CrossRef]
Figure 1. Three key limitations of ZIS-based photocatalysts for HER application.
Figure 1. Three key limitations of ZIS-based photocatalysts for HER application.
Coatings 15 01061 g001
Figure 2. The methods for the preparation of the VS-ZIS photocatalysts.
Figure 2. The methods for the preparation of the VS-ZIS photocatalysts.
Coatings 15 01061 g002
Figure 3. (a) Schematic representation of ZIS with oxygen atom doping and VS obtained by calcination under air atmosphere for overall water splitting under visible-light irradiation; (b) XRD patterns, (c) S 2p XPS spectrum, (d) EPR spectra, (e) UV-vis absorption and (f) PL spectra of the initial ZIS and the ZIS calcined at 350 °C for different times; (g) photocatalytic overall water splitting performances (50 mg; λ ≥ 420 nm) of the as-prepared samples, and (h) wavelength-dependent AQE plots of the ZIS-350-4h sample for overall water splitting. Reprinted with permission from Ref. [56]. Copyright (2023), Royal Society of Chemistry. (i) Schematic representation of rhombohedral ZIS nanosheets enriched with VS obtained by calcination under N2 atmosphere for overall water splitting under visible-light irradiation; (j) XRD patterns, (k) S 2p XPS spectra, (l) EPR spectra, (m) UV-vis absorption and (n) TRPL spectra of the initial ZIS and ZIS-800 enriched with VS; (o) photocatalytic overall water splitting performances (50 mg, λ ≥ 420 nm) of the as-prepared samples and (p) wavelength-dependent AQE plots of overall water splitting spectra of the ZIS-800 sample enriched with VS for overall water splitting. Reprinted with permission from Ref. [55]. Copyright (2022), American Chemical Society.
Figure 3. (a) Schematic representation of ZIS with oxygen atom doping and VS obtained by calcination under air atmosphere for overall water splitting under visible-light irradiation; (b) XRD patterns, (c) S 2p XPS spectrum, (d) EPR spectra, (e) UV-vis absorption and (f) PL spectra of the initial ZIS and the ZIS calcined at 350 °C for different times; (g) photocatalytic overall water splitting performances (50 mg; λ ≥ 420 nm) of the as-prepared samples, and (h) wavelength-dependent AQE plots of the ZIS-350-4h sample for overall water splitting. Reprinted with permission from Ref. [56]. Copyright (2023), Royal Society of Chemistry. (i) Schematic representation of rhombohedral ZIS nanosheets enriched with VS obtained by calcination under N2 atmosphere for overall water splitting under visible-light irradiation; (j) XRD patterns, (k) S 2p XPS spectra, (l) EPR spectra, (m) UV-vis absorption and (n) TRPL spectra of the initial ZIS and ZIS-800 enriched with VS; (o) photocatalytic overall water splitting performances (50 mg, λ ≥ 420 nm) of the as-prepared samples and (p) wavelength-dependent AQE plots of overall water splitting spectra of the ZIS-800 sample enriched with VS for overall water splitting. Reprinted with permission from Ref. [55]. Copyright (2022), American Chemical Society.
Coatings 15 01061 g003
Figure 4. (a) Schematic diagram of the synergistic effect of S-vacancy regulation and MXene co-catalytic effect in ZIS/MXene photocatalyst; (b) ZIS models with poor, moderate, and rich contents of VS and without VS; (c) calculated Gibbs free energies of hydrogen adsorption, (d) work functions, (e) S 2p XPS spectra, (f) EPR spectra, and (g) PHER performances for ZIS and ZIS-X with different levels of VS (P, M, and R represent poor, moderate, and rich contents of VS, respectively). Reprinted with permission from Ref. [61]. Copyright (2024), John Wiley & Sons Inc.
Figure 4. (a) Schematic diagram of the synergistic effect of S-vacancy regulation and MXene co-catalytic effect in ZIS/MXene photocatalyst; (b) ZIS models with poor, moderate, and rich contents of VS and without VS; (c) calculated Gibbs free energies of hydrogen adsorption, (d) work functions, (e) S 2p XPS spectra, (f) EPR spectra, and (g) PHER performances for ZIS and ZIS-X with different levels of VS (P, M, and R represent poor, moderate, and rich contents of VS, respectively). Reprinted with permission from Ref. [61]. Copyright (2024), John Wiley & Sons Inc.
Coatings 15 01061 g004
Figure 5. (a) Schematic diagram of synthesis process of ZIS with regulated surface VS; (b,c) PHER rate of ZIS and VS-ZIS under visible light; (d) sketch, and (e) EPR spectra for ZIS with regulated surface VS; (f) HER cycling test of ZIS-2 under visible light; (g) wavelength-dependent AQE of ZIS and ZIS-2. Reprinted with permission from Ref. [65]. Copyright (2023), American Chemical Society. (h) Schematic illustration on photocatalysis mechanism for water splitting; (i) escape energy of atoms in different positions, and simulated configuration of hexagonal ZIS (001) facet; (j,k) calculated DOS of pristine ZIS and VS-ZIS. (l,m) PHER rate of ZIS, different VS-ZIS under visible light. Reprinted with permission from Ref. [66]. Copyright (2021), Springer Nature.
Figure 5. (a) Schematic diagram of synthesis process of ZIS with regulated surface VS; (b,c) PHER rate of ZIS and VS-ZIS under visible light; (d) sketch, and (e) EPR spectra for ZIS with regulated surface VS; (f) HER cycling test of ZIS-2 under visible light; (g) wavelength-dependent AQE of ZIS and ZIS-2. Reprinted with permission from Ref. [65]. Copyright (2023), American Chemical Society. (h) Schematic illustration on photocatalysis mechanism for water splitting; (i) escape energy of atoms in different positions, and simulated configuration of hexagonal ZIS (001) facet; (j,k) calculated DOS of pristine ZIS and VS-ZIS. (l,m) PHER rate of ZIS, different VS-ZIS under visible light. Reprinted with permission from Ref. [66]. Copyright (2021), Springer Nature.
Coatings 15 01061 g005
Figure 6. (a) Schematic illustration of the fabrication procedure of VZn-ZIS; (b) Density of state (DOS) (upper: ZIS, lower: VZn-ZIS) and distribution of partial charge density near the CBM of ZIS and Zv-ZIS; (c) Gibbs free energy of hydrogen adsorption and (d) time-dependent HER activities of ZIS and VZn-ZIS; (e) schematic diagram of Zn vacancy promoted photocatalytic HER. Reprinted with permission from Ref. [72]. Copyright (2023), Elsevier. (f) Schematic illustration of the fabrication procedure of VZn-ZIS; (g) XRD patterns, (h) band structures, and (i) PHER performances of ZIS and VZn-ZIS samples. Reprinted with permission from Ref. [73]. Copyright (2025), Elsevier. (j) Schematic illustration of the fabrication procedure of 1.0% Pt/VZn-ZIS, (k) wavelength dependence of AQE for photo-reforming 4-MBA to HER over 1.0% Pt/VZn-ZIS, (l) free energy paths for HER. Reprinted with permission from Ref. [74]. Copyright (2023), John Wiley & Sons Inc.
Figure 6. (a) Schematic illustration of the fabrication procedure of VZn-ZIS; (b) Density of state (DOS) (upper: ZIS, lower: VZn-ZIS) and distribution of partial charge density near the CBM of ZIS and Zv-ZIS; (c) Gibbs free energy of hydrogen adsorption and (d) time-dependent HER activities of ZIS and VZn-ZIS; (e) schematic diagram of Zn vacancy promoted photocatalytic HER. Reprinted with permission from Ref. [72]. Copyright (2023), Elsevier. (f) Schematic illustration of the fabrication procedure of VZn-ZIS; (g) XRD patterns, (h) band structures, and (i) PHER performances of ZIS and VZn-ZIS samples. Reprinted with permission from Ref. [73]. Copyright (2025), Elsevier. (j) Schematic illustration of the fabrication procedure of 1.0% Pt/VZn-ZIS, (k) wavelength dependence of AQE for photo-reforming 4-MBA to HER over 1.0% Pt/VZn-ZIS, (l) free energy paths for HER. Reprinted with permission from Ref. [74]. Copyright (2023), John Wiley & Sons Inc.
Coatings 15 01061 g006
Figure 7. (a) Schematic illustration of the poor and rich VIn-ZIS photocatalysts for CO2 reduction and ACTEM images of VIn-rich-ZIS; (b) XRD patterns and (c) EPR spectra of VIn-poor ZIS and VIn-rich ZIS samples. Reprinted with permission from Ref. [76]. Copyright (2022), American Chemical Society. (d) Schematic illustration of the preparation process of ZIS-x catalysts; (e) SEM images of ZIS; (f) EPR spectra of ZIS-x catalysts; (g) the calculated charge density distribution of ZIS and ZIS-5. Inv and Znv represent the In and Zn defect sites, respectively. Reprinted with permission from Ref. [77]. Copyright (2025), American Chemical Society. (h) Mechanism of dual vacancy-engineered ZIS nanosheets; (i) PHER performances over ZIS-x catalysts under AM 1.5G illumination; (j) defective ZIS bulk DFT calculation structures and (k) piezoelectric voltage distribution within the 3D geometry for pristine, 2S, 5S + 2In, and 10S + 4In vacancies. Reprinted with permission from Ref. [78]. Copyright (2024), John Wiley & Sons Inc.
Figure 7. (a) Schematic illustration of the poor and rich VIn-ZIS photocatalysts for CO2 reduction and ACTEM images of VIn-rich-ZIS; (b) XRD patterns and (c) EPR spectra of VIn-poor ZIS and VIn-rich ZIS samples. Reprinted with permission from Ref. [76]. Copyright (2022), American Chemical Society. (d) Schematic illustration of the preparation process of ZIS-x catalysts; (e) SEM images of ZIS; (f) EPR spectra of ZIS-x catalysts; (g) the calculated charge density distribution of ZIS and ZIS-5. Inv and Znv represent the In and Zn defect sites, respectively. Reprinted with permission from Ref. [77]. Copyright (2025), American Chemical Society. (h) Mechanism of dual vacancy-engineered ZIS nanosheets; (i) PHER performances over ZIS-x catalysts under AM 1.5G illumination; (j) defective ZIS bulk DFT calculation structures and (k) piezoelectric voltage distribution within the 3D geometry for pristine, 2S, 5S + 2In, and 10S + 4In vacancies. Reprinted with permission from Ref. [78]. Copyright (2024), John Wiley & Sons Inc.
Coatings 15 01061 g007
Figure 8. (a) Schematic band structures and (b) PHER performances of undoped ZIS and transition-metal-doped ZIS. Reprinted with permission from Ref. [93]. Copyright (2011), Elsevier. (c) Schematic illustration of coupled photocatalytic reaction over Co-ZIS HNTs; (d) PHER performances and (e) PHER and benzaldehyde production rates of bare ZIS MSs and Co-ZIS HNTs. Reprinted with permission from Ref. [82]. Copyright (2022), John Wiley & Sons Inc. (f) Schematic illustration of gradient H migration for the HER on Zn facet over Cu–ZIS; (g) EPR spectra and (h) PHER performances of ZIS and xCu–ZIS (x = 1~6%). Reprinted with permission from Ref. [94]. Copyright (2021), American Chemical Society. (i) Schematic illustration of Sn2+/Sn4+-ZIS synthetic process; (j) SPV spectra and (k,l) PHER performances under visible-light irradiation (λ > 420 nm) of ZIS, Sn2+-ZIS, Sn4+-ZIS, and Sn2+/Sn4+-ZIS. Reprinted with permission from Ref. [95]. Copyright (2022), Springer Nature.
Figure 8. (a) Schematic band structures and (b) PHER performances of undoped ZIS and transition-metal-doped ZIS. Reprinted with permission from Ref. [93]. Copyright (2011), Elsevier. (c) Schematic illustration of coupled photocatalytic reaction over Co-ZIS HNTs; (d) PHER performances and (e) PHER and benzaldehyde production rates of bare ZIS MSs and Co-ZIS HNTs. Reprinted with permission from Ref. [82]. Copyright (2022), John Wiley & Sons Inc. (f) Schematic illustration of gradient H migration for the HER on Zn facet over Cu–ZIS; (g) EPR spectra and (h) PHER performances of ZIS and xCu–ZIS (x = 1~6%). Reprinted with permission from Ref. [94]. Copyright (2021), American Chemical Society. (i) Schematic illustration of Sn2+/Sn4+-ZIS synthetic process; (j) SPV spectra and (k,l) PHER performances under visible-light irradiation (λ > 420 nm) of ZIS, Sn2+-ZIS, Sn4+-ZIS, and Sn2+/Sn4+-ZIS. Reprinted with permission from Ref. [95]. Copyright (2022), Springer Nature.
Coatings 15 01061 g008
Figure 9. (a) Schematic illustration of N-doped ZIS for the PHER; (b) UV-vis absorbance spectra and (c,d) PHER performances of the N-doped ZIS under visible-light irradiation; (e) optimized structures and calculated DOS for pristine ZIS and N(4)(4′)ZIS structures. Reprinted with permission from Ref. [89]. Copyright (2023), Elsevier. (f) Schematic diagram of the energy band structures and (g) UV-vis absorbance spectra of P-doped ZIS and pure ZIS; (h,i) PHER performances of the different P-ZIS samples. Reprinted with permission from Ref. [97]. Copyright (2023), MDPI. (j) Schematic diagram of the energy band structures, (k) PL spectra, and (l) PHER performances and (m) schematic representing the higher probability of the defect-state-mediated HER of ZIS before and after doping. Reprinted with permission from Ref. [98]. Copyright (2021), American Chemical Society.
Figure 9. (a) Schematic illustration of N-doped ZIS for the PHER; (b) UV-vis absorbance spectra and (c,d) PHER performances of the N-doped ZIS under visible-light irradiation; (e) optimized structures and calculated DOS for pristine ZIS and N(4)(4′)ZIS structures. Reprinted with permission from Ref. [89]. Copyright (2023), Elsevier. (f) Schematic diagram of the energy band structures and (g) UV-vis absorbance spectra of P-doped ZIS and pure ZIS; (h,i) PHER performances of the different P-ZIS samples. Reprinted with permission from Ref. [97]. Copyright (2023), MDPI. (j) Schematic diagram of the energy band structures, (k) PL spectra, and (l) PHER performances and (m) schematic representing the higher probability of the defect-state-mediated HER of ZIS before and after doping. Reprinted with permission from Ref. [98]. Copyright (2021), American Chemical Society.
Coatings 15 01061 g009
Table 1. Comparison of HER performance of ZIS under different modification strategies.
Table 1. Comparison of HER performance of ZIS under different modification strategies.
CatalystsLight SourceScavengerHER Rate
(μmol·g−1·h−1)
Ref.
VS-ZIS
(with Pt/Cr cocatalysts)
300W Xenon lamp10 vol% acetone3.40[55]
VS-ZIS/Mo2N300W Xenon lamp15 vol% TEOA10,280[102]
VS-ZIS/Cu300W Xenon lamp
(λ > 420 nm)
0.2 M
ascorbic acid
9864.7[103]
VS-ZIS/Ni-P@C300W Xenon lamp
(λ > 400 nm)
14 vol% TEOA11,064[104]
VS-ZIS/CdS300W Xenon lamp (λ > 420 nm)10 vol% lactic aqueous16,630[105]
VS-ZIS/TpPa-1300W Xenon lamp (λ > 420 nm)0.05 M L-ascorbic acid2745[106]
VS-ZIS nanosheets300W Xenon lamp
(λ > 420 nm)
10 vol% TEOA5437[62]
VS-ZIS
(plasma-etched ZIS)
300W Xenon lampNa2S/Na2SO3
(0.35 M/0.25 M)
261.9[65]
VS-ZIS-P
(plasma-etched ZIS)
300W Xenon lampNa2S/Na2SO3
(0.35 M/0.25 M)
706[66]
VZn-ZIS nanoflowers300W Xenon lamp10 vol% TEOA21,430[73]
VZn-ZIS300W Xenon lamp10 vol% TEOA212[72]
Pt/VZn-ZIS300W Xenon lamp4-methylbenzyl alcohol6410[74]
Sn-ZIS300W Xenon lamp
(λ > 400 nm)
20 vol% TEOA62.18[96]
Pt-ZISsimulated solar light (λ > 420 nm)10 vol% TEOA17,500[107]
Pt-ZIS300W Xenon lamp15 vol% TEOA577[86]
Mn-ZIS300W Xenon lamp1% benzyl alcohol32,750[85]
Ni-ZIS300W Xenon lamp (λ > 420 nm)14 vol% TEOA8910[108]
Cu-ZIS300W Xenon lamp0.2 M ascorbic acid9864.7[94]
Pd-ZIS300W Xenon lamp0.25 M Na2SO42310[34]
Na-ZIS/CoSe2300W Xenon lamp20 vol% TEOA4525[88]
Au@Pt/ZISXenon lamp
(λ > 420 nm)
Na2S/Na2SO3
(0.35 M/0.25 M)
41,747[92]
Ag0.6Au0.4/ZIS300W Xenon lampNa2S/Na2SO3
(0.35 M/0.25 M)
54,007[91]
P-ZIS300W Xenon lamp
(420 nm < λ < 760 nm)
Na2S/Na2SO3
(0.25 M/0.35 M)
1566.6[97]
N-ZIS/(tungsten-based polyoxometalate)300W Xenon lamp
(λ>320nm)
15 vol% TEOA17,345.53[109]
N, O co-doped
ZIS
300W Xenon lampNa2S/Na2SO3
(0.35 M/0.25 M)
2254[99]
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

Hong, F.; Jing, T.; Wang, S.; He, Z. Defect Engineering of ZnIn2S4 Photocatalysts for Enhanced Hydrogen Evolution Reaction. Coatings 2025, 15, 1061. https://doi.org/10.3390/coatings15091061

AMA Style

Hong F, Jing T, Wang S, He Z. Defect Engineering of ZnIn2S4 Photocatalysts for Enhanced Hydrogen Evolution Reaction. Coatings. 2025; 15(9):1061. https://doi.org/10.3390/coatings15091061

Chicago/Turabian Style

Hong, Fangying, Tong Jing, Sen Wang, and Zuoli He. 2025. "Defect Engineering of ZnIn2S4 Photocatalysts for Enhanced Hydrogen Evolution Reaction" Coatings 15, no. 9: 1061. https://doi.org/10.3390/coatings15091061

APA Style

Hong, F., Jing, T., Wang, S., & He, Z. (2025). Defect Engineering of ZnIn2S4 Photocatalysts for Enhanced Hydrogen Evolution Reaction. Coatings, 15(9), 1061. https://doi.org/10.3390/coatings15091061

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