Jung
Wang
,
Yu
Wei
,
Bo-Hao
Zhang
,
Kai
Yang
,
Wei-Ya
Huang
and
Kang-Qiang
Lu
*
Jiangxi Provincial Key Laboratory of Functional Molecular Materials Chemistry, School of Chemistry and Chemical Engineering, Jiangxi University of Science and Technology, Ganzhou, 341000, PR China. E-mail: kqlu@jxust.edu.cn
First published on 2nd October 2025
Compared with sacrificial-agent-dependent half-reactions in photocatalytic water-splitting hydrogen production, coupling photocatalytic organic synthesis with hydrogen production markedly boosts electron–hole utilization efficiency and cuts reaction costs. In recent years, hexagonal ZnIn2S4 has been widely applied in the field of photocatalytic hydrogen production coupled with organic synthesis due to its advantages such as narrow band gap, high hydrogen evolution efficiency, good chemical stability, non-toxicity, and low cost. Herein, we present a comprehensive review of the latest progress in ZnIn2S4-based photocatalysts. We first summarize the preparation methods of ZnIn2S4 and the strategies to improve its performance, including metal doping, morphological engineering, heterostructure construction and defect engineering. Subsequently, we focus on the research progress of ZnIn2S4-based photocatalysts in hydrogen production coupling organic synthesis, including the selective conversion of alcohols, oxidative coupling of amines, thiol dehydrogenation, and biomass oxidation. Finally, the challenges and opportunities that ZnIn2S4 faces in practical application are discussed. It is expected that this review will offer insightful guidance for the rational design of semiconductor-based dual-functional photoredox reaction systems, thereby injecting impetus into the research on harvesting environmentally solar fuel production as well as the high-value-added fine chemicals.
Among numerous photocatalysts, ZnIn2S4 has been widely employed in the field of photocatalysis due to its facile synthesis, broad light absorption range, and excellent photocatalytic performance.9–12 Its primary applications in energy conversion include visible-light-driven photocatalytic water splitting for H2 production and photocatalytic CO2 reduction.13–17 As illustrated in Fig. 1, the ZnIn2S4-based photocatalyst has been extensively employed in the field of photocatalysis in recent years, with an increasing trend over successive years. ZnIn2S4 possesses suitable band positions and a unique layered structure, endowing ZnIn2S4-based photocatalysts with remarkable photocatalytic activity in H2 production coupling organic synthesis.18–20 To date, various strategies have been developed to design ZnIn2S4-based photocatalysts, including element doping, morphology control, cocatalyst modification, and heterostructure construction.21,22 These strategies effectively enhance the charge carrier separation efficiency and light absorption capacity of ZnIn2S4-based photocatalysts, thereby improving their performance in photocatalytic H2 production coupling organic synthesis. However, there persists an absence of a comprehensive review that systematically summarizes the applications of ZnIn2S4 in photocatalytic hydrogen production coupled with various organic synthesis reactions.
![]() | ||
| Fig. 1 The illustration shows the number of papers published each year on the topic “photocatalysts, & ZnIn2S4” since 2018. | ||
Therefore, this review comprehensively summarizes the structure of ZnIn2S4 and introduces several common synthetic methods for ZnIn2S4-based photocatalysts. Additionally, we emphasize the modification and optimization of ZnIn2S4-based composites and summarize their applications in photocatalytic H2 production coupling organic value added. Finally, we discuss the current challenges faced by ZnIn2S4-based composites in coupling reactions and propose core issues that need to be addressed in the future. We anticipate that this review will deepen the understanding of photocatalytic H2 production coupling organic synthesis and facilitate the design and development of advanced photocatalysts.
:
2) and twice the amount of thiourea in water, methanol and ethylene glycol, and thermos-reacted at 140 °C in the solvent for 72 h. Different solvents resulted in different morphologies. Aqueous solvent corresponds to the nanosheet structure, methanol solvent corresponds to the nanoporous structure, and ethylene glycol corresponds to random nanoparticles without a specific morphology.
In 2024, Wang et al.42 synthesized ZnIn2S4 with sulfur vacancies (ZnIn2S4-Vs) by solvothermal binding annealing (Fig. 3a). Workers obtain uniformly assembled ZnIn2S4 ultrathin nanosheets at low temperatures and short reaction times through a simple and safe water bath reaction. Then, ZnIn2S4 with S vacancies is obtained by annealing ZnIn2S4. As illustrated in Fig. 3b, Wang et al.43 prepared a ZnIn2S4@Ni(OH)2/NiO Z-scheme heterostructure photocatalyst with defects by using hydrothermal synthesis combined with low temperature roasting. In 2025, Li et al.44 synthesized NiCo2O4@ZnIn2S4 composites in a two-step process (Fig. 3c). First, NiCo2O4 hollow nanoprisms were prepared by co-precipitation and the Kirkendall effect. This is then incorporated into the steps of conventional solvothermal synthesis of ZnIn2S4.
![]() | ||
| Fig. 3 (a) Fabrication process of Vs-ZnIn2S4. Reproduced with permission from ref. 42, copyright 2024, Wiley-VCH. (b) The schematic illustration of the synthesis process of ZnIn2S4, ZnIn2S4@NOH and ZnIn2S4@NOH/NO photocatalysts. Reproduced with permission from ref. 43, copyright 2024, Elsevier. (c) Schematic illustration for the fabrication of NiCo2O4@ZnIn2S4. Reproduced with permission from ref. 44, copyright 2025, Elsevier. | ||
![]() | ||
| Fig. 4 (a) Photocatalytic degradation mechanism for pollutants over CoSx/Ag/Ag2S/ZnIn2S4 composites under visible light. Reproduced with permission from ref. 65, copyright 2022, Elsevier. (b) Schematic illustration of charge transfer and the H2 evolution mechanism for the Ag-Pd/ZnIn2S4 photocatalyst. Reproduced with permission from ref. 66, copyright 2023, Elsevier. (c) The schematic of the reaction mechanism of ZnIn2S4 and 2-Ni/ZnIn2S4 for H2 evolution. Reproduced with permission from ref. 68, copyright 2022, Elsevier. (d) Schematic diagram of the photocatalytic mechanism for Bi-Bi2MoO6/ZnIn2S4 under visible light illumination and (e) photocatalytic H2 producing performance of the prepared samples. Reproduced with permission from ref. 69, copyright 2023, Elsevier. | ||
Due to the high cost of precious metals and the scarcity of resources, their practical application is somewhat limited, so workers began to engage in the adulteration of non-noble metals. Lin et al.67 prepared the Ni/ZnIn2S4 composite photocatalyst by doping with non-noble metal Ni, and proved that Ni doping effectively inhibits electron–hole recombination. Therefore, the activity of the composite photocatalyst Ni/ZnIn2S4 in the reaction of hydrogen production coupled with BA oxidation was significantly improved. Meanwhile, Chen et al.68 employed an in situ photodeposition strategy to anchor Ni clusters on the ZnIn2S4 surface, which yielded a 10.6-fold enhancement in hydrogen evolution activity. The underlying reaction mechanism is schematically illustrated in Fig. 4c. Ni-doped photocatalysts have not only overcome the disadvantages of expensive precious metals in the past, but also provide inspiration for the doping of other monometals and their complexes. Furthermore, Geng et al.69 utilized Bi as a noble metal substitute to enhance the active sites of ZnIn2S4 and induce SPR effects (Fig. 4d). As shown in Fig. 4e, this composite catalyst exhibited superior photocatalytic activity in hydrogen production compared to its Pt-doped ZnIn2S4 counterparts.
Moreover, Liu et al.73 reported a spatial decoupling strategy assembly with the cryogenic solvent method forming a high-quality tight heterojunction structure CoP-ZnIn2S4 composite photocatalyst (Fig. 5a). As shown in Fig. 5b–d, the CoP-ZnIn2S4 composite photocatalyst was used as a model to construct unique CoP coaxial nanorods. CoP as the core is the active site for hydrogen production, and ZnIn2S4 as the shell is the oxidative coupling site of amine to imine. This structure has a larger areal density, which provides more reaction sites, and effectively ensures the effective isolation of the active sites of H2 and imine, and improves its product yield and selectivity. Fig. 5e illustrates the synthesis of ZnIn2S4 nanosheets grown by in situ confinement on a porous hollow carbon sphere (PHCS) backbone through the hydrothermal method.74 SEM images of the carbon spheres show a diameter of approximately 250 nm and also show a hollow structure and numerous micropores (Fig. 5f). After the hydrothermal treatment, the ZnIn2S4/C heterostructure composites retained the original morphology of the carbon framework, and a small number of ZnIn2S4 nanosheets were grown inside/outside the frame (Fig. 5g), and the high-resolution TEM images showed the interfacial composition of the composites (Fig. 5h). It is the exploration of the morphology of ZnIn2S4 that has led to the emergence of more excellent ZnIn2S4 composites with outstanding structural and catalytic properties.
![]() | ||
| Fig. 5 (a) Schematic illustration for the synthetic process of CoP@ZnIn2S4, (b) TEM, (c) SEM, and (d) high-resolution TEM images of CoP@ZnIn2S4. Reproduced with permission from ref. 73, copyright 2023, American Chemical Society. (e) Fabrication scheme of the hierarchical hollow structural ZnIn2S4/C composite, (f) SEM image of PHCSs, and (g) TEM image and (h) HRTEM images of ZnIn2S4/C. Reproduced with permission from ref. 74, copyright 2025, Elsevier. | ||
For example, Fig. 6a shows a bifunctional 2D/2D ZnIn2S4/CeO2 photocatalyst with a stepped heterojunction which was successfully synthesized by Zhang et al.78 Jiang et al.79 also synthesized the CeO2/ZnIn2S4 layer hybrids by the in situ solvothermal method. Given that the chemical properties of the sulfur (S) atom are similar to those of the oxygen (O) atom, the S2− is able to fill the oxygen vacancies in CeO2. And a novel bifunctional 2D/2D CeO2/ZnIn2S4 photocatalyst was successfully developed. The catalyst has a step-like heterostructure and can simultaneously utilize the electron–hole pair generated by photoexcitation to efficiently produce H2 and oxidize furfural in a single photooxidation–reduction reaction system. Furthermore, in the EPR spectroscopic analysis, ZnIn2S4 nanospheres and CeO2/ZnIn2S4 show an obvious DMPO-˙O2− signal, and show weaker signals compared with the EPR spectra of cerium oxide nanorods (Fig. 6b). From another perspective, CeO2 and CeO2/ZnIn2S4 composites exhibit more pronounced significant DMPO-˙OH signals in the EPR spectra, characterized by a typical peak intensity ratio of 1
:
2
:
2
:
1, whereas no such peaks are observed for ZnIn2S4 nanospheres (Fig. 6c). In addition, the results of band structure and EPR spectra indicate that the CeO2/ZnIn2S4 system successfully constructs a Z-type structure (Fig. 6d). Li et al.80 synthesized the WO3/ZnIn2S4 heterojunction structure and proposed a reasonable photoredox catalytic reaction mechanism (Fig. 6e).
![]() | ||
| Fig. 6 (a) A reasonable catalytic mechanism for the cooperative photoredox reaction including H2 production and FOL conversion over the 2D/2D ZnIn2S4/CeO2 heterojunction. Reproduced with permission from ref. 78, copyright 2023, Elsevier. (b) EPR spectra of DMPO-˙O2− and (c) DMPO-˙OH over CeO2, (d) ZnIn2S4 and CeO2/ZnIn2S4 under solar exposure, and charge carrier flow mechanisms over type II and direct Z-scheme heterojunctions for the CeO2/ZnIn2S4 composite. Reproduced with permission from ref. 79, copyright 2020, Elsevier. (e) Probable mechanism for photoredox-catalyzed integrated reactions including BA oxidation and H2 generation over WO3/ZnIn2S4 under UV-vis light illumination. Reproduced with permission from ref. 80, copyright 2022, American Chemical Society. | ||
Recently, Yang et al.85 successfully synthesized ZnIn2S4 nanosheets with sulfur vacancies by tuning the concentration of TAA (Fig. 7a). Notably, this strategy not only enabled precise modulation of the ZnIn2S4 band structure but also markedly augmented its piezoelectric effect, thereby laying a structural foundation for enhanced photocatalytic performance. Furthermore, Yue et al.86 engineered a built-in strong electric field by introducing S vacancies into ZnIn2S4 and integrating with NiSe. This electrostatic field effectively mitigated the coulombic attraction between electron–hole pairs to suppress their recombination, thereby achieving highly active photocatalytic hydrogen evolution. The corresponding synthesis route is illustrated in Fig. 7b. The functionality of this built-in electric field is directly validated by the electrostatic potential distributions presented in Fig. 7c and d, where a significant enhancement in the potential difference was observed after the introduction of S vacancies. Fig. 7e depicts the DFT-calculated crystal structure model of ZnIn2S4-Sv along the [001] direction.
![]() | ||
| Fig. 7 (a) The schematic illustrates the synthesis of S-vacancy-engineered ZnIn2S4 nanosheets. Reproduced with permission from ref. 85, copyright 2025, Wiley-VCH. (b) Schematic illustration of the synthetic process for the ZnIn2S4-NiSe, and NiSe/ZnIn2S4-Sv samples, (c) DFT calculations yielded the electrostatic potential diagrams along the [001] direction for ZnIn2S4 and (d) ZnIn2S4-Sv, and (e) the crystal structure diagram along the [001] direction for ZnIn2S4-Sv. Reproduced with permission from ref. 86, copyright 2024, Wiley-VCH. | ||
In the design of ZnIn2S4-based composite photocatalysts, the introduction of appropriate co-catalysts is of great significance for enhancing charge separation efficiency and activating reaction sites—two core factors that determine the performance of selective alcohol oxidation. Recently, Peng et al.53 described the PPy@ZnIn2S4 composite materials which involve the direct growth of ultra-thin ZnIn2S4 nanosheets on PPy nanotubes (Fig. 8a). This composite material exhibits excellent photocatalytic activity capable of generating H2 through the photocatalytic reduction reaction of water and producing 1,4-phenylacetaldehyde (PAD) through the selective oxidation reaction of 1,4-benzenedimethanol (BDM). And the introduction of PPy is mainly to improve the activity of the photoreaction by increasing the sulfur vacancies of ZnIn2S4. As shown in the figures, the Gibbs free energies of all possible reactions during the BDM oxidation reaction were calculated for ZnIn2S4 and ZnIn2S4-Sv (Fig. 8b and c). At the same time, path 4 in the free energy diagram is identified as the best reaction path of BDM oxidation. As shown in Fig. 8d, the BDM oxidation reaction on the ZnIn2S4-Sv has a lower free energy than that on ZnIn2S4 (0.94 eV and 1.06 eV, respectively), indicating that ZnIn2S4-Sv exhibits superior catalytic activity in this oxidation reaction. Therefore, the introduction of PPy can improve the catalytic effect of the BDM oxidation reaction. As shown in Fig. 8e, it is proved that the introduction of PPy into ZnIn2S4 can also improve the hydrogen production activity. According to the above analysis, the introduction of PPy into ZnIn2S4 can significantly improve the performance of the BDM oxidation reaction and HER process.
![]() | ||
| Fig. 8 (a) Schematic preparation of PPy@ZnIn2S4, (b) the free energy diagrams of 4 pathways in ZnIn2S4 and (c) ZnIn2S4-Sv, (d) the free energy profiles with the lowest energy barrier of ZnIn2S4 and ZnIn2S4-Sv are presented by blue and red lines, respectively, and (e) the free energy profiles for hydrogen adsorption at different active sites of ZnIn2S4 and ZnIn2S4-Sv. Reproduced with permission from ref. 53, copyright 2022, Elsevier. (f) Schematic illustration of the Z-scheme transfer path of photogenerated carriers between HPM and ZnIn2S4 heterojunctions. HPB refers to the reduction state of an HPM derived from the formation of a built-in electric field. Reproduced with permission from ref. 90, copyright 2022, Elsevier. | ||
Moreover, for ZnIn2S4-based photocatalytic systems, constructing heterojunctions is also a crucial performance optimization strategy—this structure not only facilitates the efficient separation of photogenerated electron–hole pairs (a core bottleneck in photocatalysis) but also regulates the reaction path to precisely match the redox potential requirements of alcohol oxidation coupling with H2 evolution. Huang et al.90 reported an efficient aromatic alcohol oxidation with simultaneous H2 evolution under aqueous conditions, achieved in a polyoxometalate (POM)-incorporated ZnIn2S4 dual-functional photocatalytic system. And the schematic illustration of the Z-scheme path between HPM and ZnIn2S4 heterojunctions is shown in Fig. 8f.
For ZnIn2S4-based photocatalysts, atomic-level doping modification has emerged as a powerful strategy to overcome performance bottlenecks—it not only constructs well-defined atomic active sites to strengthen the adsorption and activation of reactants but also regulates electronic effects to optimize the dynamics of photogenerated charge transfer, thereby synergistically enhancing catalytic activity for coupled reactions such as H2 evolution coupling with aromatic alcohol oxidation. In 2025, Zhou et al.91 employed an approach by doping non-magnetic molybdenum (Mo) into ZnIn2S4 materials, while the spin polarization effect is realized and S22−–Mo–S–In active sites are formed at the atomic level. The synergistic coupling of spin polarization and atomic active sites is the key driving force for the significant enhancement of catalytic performance. As shown in Fig. 9g and h, the optimized Mo/ZnIn2S4 showed a catalytic activity of 172.26 mmol g−1 h−1 and 161.46 mmol g−1 h−1 in the production of hydrogen and benzaldehyde, respectively, which was nearly 10 times higher than that of ZnIn2S4. Moreover, further improvements in hydrogen and BAD production can be achieved by the external magnetic field generated by the permanent magnet without any additional energy input. Furthermore, the microscopic coordination environment of the photocatalyst was characterized by X-ray absorption near edge structure (XANES) and extended X-ray absorption fine structure (EXAFS) spectroscopy techniques (Fig. 9a–f). It was confirmed that Zn exists mainly in the form of Zn-S coordination and Mo is mainly coordinated with S atoms.
![]() | ||
| Fig. 9 (a) XANES spectra of the Zn K-edge for Zn foil, ZnO, ZnIn2S4 and Mo/ZnIn2S4, (b) the Fourier transform of k2− weighted EXAFS spectra at the Zn k-edge of Zn foil, ZnO, ZnIn2S4 and 2Mo/ZnIn2S4, (d) XANES spectra of the Mo K-edge for Mo foil, MoS2 and Mo/ZnIn2S4, (e) the Fourier transform of k2− weighted EXAFS spectra at the Mo K-edge of Mo foil, MoS2 and Mo/ZnIn2S4, (c) wavelet transform analysis of the k2− weighted Mo K-edge EXAFS signals of Mo foil, MoS2 and (f) Mo/ZnIn2S4. Reproduced with permission from ref. 91, copyright 2025, Elsevier. | ||
Meanwhile, Jia et al.92 put forward a new type of 2D/3D ZnIn2S4@Ni1/UiO-66-NH2 heterojunction, and the spatial separation and ordered distribution of oxidation–reduction sites were realized for photocatalytic H2 production and BA value addition. The spatial separation and directional electron transfer of redox sites were fully realized. EXAFS spectroscopy analysis illustrates the local environment of Ni to detect the geometrical and electronic structure of the Ni site. Fig. 10a shows a composite diagram of ZnIn2S4@Ni1/UiO-66-NH2. In addition, Fig. 10b proves that Ni exists in the form of Ni2+. The wavelet transform of the composites shows that this verifies the ligation environment of Ni-O and the single-atom properties of Ni (Fig. 10c). The Fourier transform-EXAFS spectra showed that there were no peaks associated with Ni–Ni bonds, indicating that the Ni2+ sites were all dispersed (Fig. 10d). The curve fitting of the EXAFS data also indicates that the Ni2+ site has a coordinated environment. As shown in Fig. 10e, the best fitting results of the first shell show that Ni is mainly coordinated with oxygen atoms.
![]() | ||
| Fig. 10 (a) Schematic illustration of the synthetic process of ZnIn2S4@Ni1/UN nanocomposites, (b–d) Ni K-edge XANES and FT-EXAFS spectra of Ni1/UN-6, Ni foil, and Ni-Pc, and (e) EXAFS fitting of Ni1/UN-6 (inset: supposed structure of a Ni single-atom anchored on a Zr6-oxo cluster). Reproduced with permission from ref. 92, copyright 2024, Elsevier. | ||
To gain insights into the research progress of alcohols being selectively converted to aldehydes along with hydrogen production using ZnIn2S4-based photocatalysts, Table 1 summarizes recent relevant studies. It presents details such as the specific photocatalyst, synthesis method, target product, light source employed, and key performance indicators.
| Photocatalyst | Synthesis method | Product | Light source | Efficiency (mmol g−1 h−1) or selectivity | Ref. |
|---|---|---|---|---|---|
| a The E and S in the table represent the yield and selectivity of the corresponding organic compounds. | |||||
| POM/ZIS | Solvothermal | Benzaldehyde | Xenon lamp (λ ≥ 420 nm 5 h) | H2: 10.60 | 90 |
| S: 100% | |||||
| PPy/ZIS | Solvothermal | 1,4-Phenyldicarboxaldehyde | Xenon lamp (λ ≥ 420 nm 6 h) | H2: 0.73 | 53 |
| S: 98.90% | |||||
| Bi/ZIS | One-pot | Benzaldehyde | Xenon lamp (λ ≥ 420 nm 6 h) | H2: 3.66 | 93 |
| E: 1.02 | |||||
| P-MoS2-ZIS | Solvothermal | Benzaldehyde | Sunlight (λ: 320–780 nm) | H2: 31.30 | 94 |
| E: 31.40 | |||||
| CeO2/ZIS | One-pot | Benzaldehyde | Xenon lamp (AM 1.5 3 h) | H2: 1.50 | 79 |
| E: 0.66 | |||||
| Au/ZIS | Solvothermal& in suit deposition | Benzaldehyde | Xenon lamp (λ ≥ 420 nm 4 h) | H2: 0.33 | 64 |
| E: 0.35 | |||||
| ZIS NS | One-pot | Benzaldehyde | Xenon lamp (AM 1.5 4 h) | H2: 22.20 | 72 |
| E: 0.33 | |||||
| CoP/ZIS | Hydrothermal & solvothermal | Pyruvic acid | Xenon lamp (λ ≥ 420 nm 3 h) | H2: 0.23 | 95 |
| S: 93.90% | |||||
| MoS2/ZIS | Solvothermal & photodeposition | Benzaldehyde | Xenon lamp (λ ≥ 420 nm 5 h) | H2: 3.88 | 54 |
| E: 3.69 | |||||
| Ni/V ZIS | In situ photodeposition | Benzaldehyde | Xenon lamp (λ: 400–800 nm 2 h) | H2: 0.82 | 85 |
| E: 0.74 | |||||
| NiP/ZIS | Solvothermal | Pyruvic acid | Xenon lamp (λ: 345–385 nm 3 h) | H2: 0.38 | 20 |
| S: 97.80% | |||||
| CoCuP/ZIS | Grinding sintering | Acetone | Xenon lamp (λ: 345 nm, 420 nm, 500 nm 3 h) | H2: 26.80 | 96 |
| E: 17.80 | |||||
| CdS/ZIS | Solvothermal | Benzaldehyde | Xenon lamp (λ: 400 nm 6 h) | H2: 12.20 | 97 |
| E: 12.10 | |||||
![]() | ||
| Fig. 11 (a) Illustration of the preparation of a sandwich-like hollowed Pd@TiO2@ZnIn2S4 nanobox catalyst. Reproduced with permission from ref. 99, copyright 2022, Wiley-VCH. (b) The calculated energy profile for hydrogen production on BaTiO3, ZnIn2S4 and BaTiO3-ZnIn2S4. Reproduced with permission from ref. 100, copyright 2023, Elsevier. (c) Schematic illustration of the dual-functional photocatalytic reaction mechanism for hydrogen evolution and BA oxidation over Pt@UiO-66-NH2@ZnIn2S4. Reproduced with permission from ref. 101, copyright 2023, Royal Society of Chemistry. | ||
Furthermore, to enhance both stability and activity, ZnIn2S4 was simultaneously engineered on two fronts: hetero-atom doping was used to re-tune its band structure, and a chemically inert oxide over-layer provided robust surface shielding. The two tactics operate in concert, efficiently suppressing photo-corrosion and locking the lattice integrity over prolonged illumination. Building on this stabilised matrix, Wang et al. further assembled a layered Z-scheme BaTiO3@ZnIn2S4 heterojunction that integrates piezo-electricity with photochemistry. Under mechanical vibration the built-in piezo-field triggers a “domino” displacement of charges, accelerating Z-path carrier separation and delivering photogenerated electrons and holes to their respective reaction sites with minimal recombination. Exploiting this synergy of piezo-potential and Z-scheme architecture, the composite co-produces 5.59 mmol g−1 of C–N coupled products and 8.04 mmol g−1 of H2 in a single run, demonstrating that mechanical energy can be seamlessly married to solar energy to boost photocatalytic efficiency. Mechanistic investigation further unveiled a consecutive dehydrogenation C–N coupling sequence for benzylamine, clarifying how the dual-energy input governs reaction selectivity as shown in Fig. 11b.
Similarly, the integration of heterojunction construction and morphological optimization has also been proven to exert a positive effect on promoting the photocatalytic C–N coupling reaction with synchronous H2 evolution—a trend consistent with the aforementioned performance enhancement strategies for ZnIn2S4-based systems. The Pt@UiO-66-NH2@ZnIn2S4 heterojunction with an egg yolk–shell structure was successfully prepared and used as a bifunctional photocatalyst for the C–N coupling reaction with H2 production by Hou et al.101 In this coupling system, ZnIn2S4 acts as a hole collector, and is able to capture and release holes from Pt@UiO-66-NH2, thereby accelerating the C–N coupling of benzylamine, while Pt acts as an electron collector and is also the active site for the reduction of protons to H2. The composite photocatalysts show 4 and 16 times higher catalytic activities than that of pure ZnIn2S4 and UiO-66-NH2, respectively. In addition, the well-designed and fabricated layered Pt@UiO-66-NH2@ZnIn2S4 heterostructure enables the simultaneous generation of H2 and C–N coupling reaction without the use of sacrificial agents. For the first time, the active site separation of Pt and ZnIn2S4 was successfully achieved in the interior and surface of UiO-66-NH2, which significantly enhanced the separation efficiency and extended the lifetime of the charges. A bifunctional photocatalytic reaction mechanism shown in Fig. 11c involving the oxidative coupling of aromatic amines generated by H2 binding on Pt@UiO-66-NH2@ZnIn2S4 is proposed.
In addition, the sulfur–vacancy (Sv) density in ZnIn2S4 can be precisely tuned by synergistic rare-earth-ion doping and in situ cocatalyst decoration, a strategy that concurrently accelerates H2 evolution, promotes C–N coupling, and steers the selectivity toward imines. Xie et al.102 synthesized a nanoflower-like ZnIn2S4 by dissolution heat synthesis. The Sv generated during the synthesis of ZnIn2S4 was doped with rare earth ion Er, and the NiS cocatalyst was grown on the SV by in situ photodeposition. EPR tests were performed on all three samples to study sulfur vacancies. As shown in Fig. 12d, the ZnIn2S4, 2% Er3+–ZnIn2S4, and 2% Er3+–ZnIn2S4/NiS (0.5%) samples exhibit a pattern of EPR characteristic signals at g = 2.001. This demonstrates that the number of S vacancies decreases with the increase in Er3+ doping and cocatalyst loading. Subsequently, the complex structures of ZnIn2S4, 2% Er3+–ZnIn2S4, and 2% Er3+–ZnIn2S4/NiS (0.5%) were investigated using Vienna ab initio simulation package-based density functional theory calculations (Fig. 12a and c). Since In is more electronegative than Zn, bond cleavage mainly occurs in zinc–sulfur bonds, which allows Er to replace Zn sites, thus forming Er–S bonds and reducing sulfur vacancies. After the cocatalyst NiS was loaded, the cocatalyst NiS was grown on the S vacancy on the surface of ZnIn2S4 by in situ photodeposition, which further reduced the Sv and made the cocatalyst load more stable. The obtained Er3+–ZnIn2S4/NiS photocatalyst significantly increased the hydrogen generation and C–N coupling productivity of 11.8 mmol g−1 h−1 and 10.2 mmol g−1 h−1, respectively, while the hydrogen generation rate of the original ZnIn2S4 was 1.2 mmol g−1 h−1 and the benzylamine coupling activity was 0.7 mmol g−1 h−1. Moreover, the Electron Localization Function (ELF) also indirectly confirmed the reduction of Sv, which stabilized the composite structure and further promoted the improvement of activity (Fig. 12e). In the 2% Er3+–ZnIn2S4/NiS (0.5%) system, imine was the main product, indicating that the C–C coupling reaction was significantly inhibited, thereby improving the selectivity of imine. This improvement may be attributed to the high efficiency of the NiS-assisted process. The 2% Er3+–ZnIn2S4/NiS (0.5%) system significantly inhibited the C–C coupling reaction, and the selectivity of imine was improved and became the main reaction product. This optimization may be due to the efficient synergistic effect of NiS.
![]() | ||
| Fig. 12 (a) Crystal structure model of ZnIn2S4, (b) Er3+–ZnIn2S4 and (c) 2% Er3+–ZnIn2S4/NiS (0.5%), (d) the EPR spectra of ZnIn2S4, 2% Er3+–ZnIn2S4 and 2% Er3+–ZnIn2S4/NiS (0.5%), and (e) ELF spectra of ZnIn2S4, 2% Er3+–ZnIn2S4 and 2% Er3+–ZnIn2S4/NiS (0.5%). Reproduced with permission from ref. 102, copyright 2025, Elsevier. (f) Time-yield plots of NBI and DBA, and NBI selectivity of ZnIn2S4 and (g) CoP@ZnIn2S4. Reproduced with permission from ref. 73, copyright 2023, American Chemical Society. | ||
In contrast, CoP@ZnIn2S4 composites were more selective for C–C coupling products, with the NBI selectivity decreasing to 20% after a 5 hour reaction on the original ZnIn2S4 (Fig. 12f).73 In contrast, the NBI generation process on CoP@ZnIn2S4 coaxial nanorods was efficient and stable, and the selectivity of NBI was maintained at about 100% (Fig. 12g). It is evident that the main products of ZnIn2S4 and CoP@ZnIn2S4 are DBA and NBI, respectively. This also provides researchers with more ideas for screening different coupling products of BA.
To gain insights into the research progress of oxidative coupling of amines and hydrogen production via ZnIn2S4-based photocatalysts, Table 2 summarizes recent relevant studies. It presents details such as the specific photocatalyst, synthesis method, target product, light source employed, and key performance indicators.
| Photocatalyst | Synthesis method | Product | Light source | Efficiency (mmol g−1 h−1) or selectivity | Ref. |
|---|---|---|---|---|---|
| a The E and S in the table represent the yield and selectivity of the corresponding organic compounds. | |||||
| BaTiO3/ZIS | Solvothermal | N-benzylbenzylamine | Sunlight (λ: 320 nm-780 nm) | H2: 8.04 | 100 |
| E: 5.59 | |||||
| Pd/ZIS | in situ icing assisted photo-reduction | N-benzylidenebenzylamine | Xenon lamp (λ ≥ 420 nm 2 h) | H2: 11.10 | 103 |
| E: 10.20 | |||||
| CuS & RuS/ZIS | in situ deposition | N-benzylbenzaldimine | Xenon lamp (λ ≥ 420 nm 6 h) | H2: 0.081 | 104 |
| S: 99.90% | |||||
| CoP/ZIS | Self-assembly | N-benzylbenzaldimine | Xenon lamp (λ ≥ 400 nm 5 h) | H2: 3.84 | 73 |
| E: 3.82 | |||||
| DZIS | Hydrothermal | N-benzylbenzaldimine | Xenon lamp (λ ≥ 420 nm 2 h) | H2: 4.65 | 105 |
| E: 7.39 | |||||
| Pt/UiO-66-NH2/ZIS | Solvothermal | N-benzylbenzaldimine | Xenon lamp (λ ≥ 420 nm 8 h) | H2: 0.85 | 101 |
| S: 78.00% | |||||
| Er3+-ZnIn2S4/NiS | Hydrothermal | N-benzylbenzaldimine | Xenon lamp (λ ≥ 420 nm 2 h) | H2: 11.80 | 102 |
| E: 11.20 | |||||
Specific cocatalyst modification strategies can significantly optimize the photocatalytic performance of semiconductor materials in thiol dehydrogenation-coupled reactions, solving the problem of low efficiency caused by carrier recombination. For example, Li et al.106 successfully synthesized the ZnIn2S4 photocatalyst decorated with PdS by hydrothermal and in situ deposition for coupling thiol dehydrogenation to high value-added disulfide and green hydrogen energy (Fig. 13a). They report the synthesis of a PdS modified ZnIn2S4 (PdS-ZnIn2S4) composite material and its application of photocatalytic thiol coupling to disulfides and hydrogen evolution under visible light. The blank ZnIn2S4 catalyst exhibited low yields of bis(4-methoxyphenyl) disulfide (5.4 μmol) and H2 (4.2 μmol), with a 4-methoxythiophenol conversion rate of only 11.1%, due to rapid carrier recombination. After modification with a PdS cocatalyst, the catalytic performance improved significantly: bis(4-methoxyphenyl) disulfide yield reached 45.6 μmol (8.4-fold increase), H2 yield 44.9 μmol (10.7-fold increase), and 4-methoxythiophenol conversion rate increased to 91.6%. The sulfur-central radical is a significant reaction intermediate in this coupling reaction. Furthermore, the PdS-ZnIn2S4 composites have been demonstrated to effectively catalyze various sulfhydryl compounds with different substituents to generate the corresponding S–S conjugates. As shown in Fig. 13b, a built-in electric field is formed between PdS and ZnIn2S4. When irradiated with visible light, the photogenerated holes in the VB of ZnIn2S4 migrate rapidly to PdS under the action of the built-in electric field. Moreover, under visible light irradiation, the following possible mechanism has been proposed to explain the coupling process of photocatalytic dehydrogenation of 4-methylthiophenol and precipitation of H2 from PdS-ZnIn2S4 composites (Fig. 13c).
![]() | ||
| Fig. 13 (a) Systematic illustration of the synthetic procedure of the PdS decorated ZnIn2S4 composites, (b) schematic diagram of the internal electric field formed between ZnIn2S4 and PdS, and (c) systematic illustration of the photoredox catalyzed mechanism for dehydrogenation coupling of 4-methoxythiophenol (4-MTP) with bis(4-methoxyphenyl) disulfide (4-MPD) and H2 evolution over the PdS-ZnIn2S4 photocatalyst. Reproduced with permission from ref. 106, copyright 2023, Elsevier. (d) UV-vis spectra of ZnIn2S4 and different wt% PtS/ZnIn2S4 and (e) proposed mechanism for photocatalytic splitting of thiols to produce disulfides and hydrogen over PtS/ZnIn2S4 under visible light irradiation. Reproduced with permission from ref. 107, copyright 2018, Elsevier. | ||
In addition, in-depth exploration of the catalytic universality and reaction mechanism of modified ZnIn2S4 composites is crucial for promoting the practical application of photocatalytic thiol conversion systems, and multi-case studies can provide comprehensive theoretical support. Zhang et al.107 reported the formation of small Pt nanoparticles with surface deposition by photoreduction [PtCl6]2– on hexagonal ZnIn2S4. The resulting Pt/ZnIn2S4 nanocomposites exhibit excellent photocatalytic thiol lysis to generate disulfide and hydrogen production under visible light. Isotope labeling was used to prove that the main source of hydrogen is thiols rather than water. As shown in Fig. 13d, the UV-vis diffuse reflectance spectroscopy of the PtS/ZnIn2S4 nanocomposites shows that although the loading of PtS does not change the bandgap width of ZnIn2S4, the composites show significantly enhanced light absorption in the visible region from 420 to 620 nm. Furthermore, a possible photocatalytic reaction mechanism was proposed to explain the process of thiol cleavage to disulfide on PtS/ZnIn2S4 nanocomposites (Fig. 13e).
![]() | ||
| Fig. 14 (a) Schematic illustration of the synthetic route of the ReS2/ZnIn2S4-Sv heterojunction for a photo-redox dual reaction system and (b) photocatalytic reaction mechanism of the ReS2/ZnIn2S4-Sv heterojunction under visible light irradiation. Reproduced with permission from ref. 111, copyright 2023, Elsevier. (c) DFF and H2 production rates and DFF selectivity over different samples and (d) the calculated adsorption free energy of H (ΔGH*) on different atoms of the Ni1/ZnIn2S4 photocatalyst. Reproduced with permission from ref. 112, copyright 2024, American Chemical Society. | ||
In addition, the construction of atomically dispersed active sites is also the core regulation method to improve the selectivity and efficiency of ZnIn2S4-based photocatalysts in biomass conversion coupling hydrogen production. Si et al.112 rationally constructed the structure of atomically dispersed Ni on ZnIn2S4 nanosheets (Ni1/ZnIn2S4), and the active site that can achieve efficient photocatalytic oxidation of 5-hydroxymethylfurfural (HMF) accompanied by H2 generation was revealed. Under visible light irradiation, Ni1/ZnIn2S4 was significantly superior to its original ZnIn2S4 form, and could catalyze the formation of 2,5-diformylfuran (DFF) with high selectivity (>97%), and achieve high yields of DFF generation (0.39 mmol g−1 h−1) and H2 generation (0.34 mmol g−1 h−1). The selective oxidation and H2 generation performance of HMF were also compared with those of other reported photocatalysts (Fig. 14c). This work provides new insights into the development of artificial photosynthesis for the upscaling of value-added chemicals from biomass through the rational construction of atomic dispersion active sites. As shown in Fig. 14d, a photocatalytic mechanism based on the selective oxidation of HMF dominated by photogenerated holes and •OH to generate DFF accompanied by H2 release was proposed.
Furthermore, to investigate the effect of adding a noble metal catalyst on the reaction, Wang et al.113 reported that a ZnIn2S4/Nb2O5 Z-type heterostructure photocatalyst was constructed by growing Nb2O5 nanospheres on the petals of ZnIn2S4 microspheres. Moreover, by comparing whether there was a Pt source added in the reaction process, the optimal conditions of the reaction were explored. As shown in Fig. 15a, the yield of H2 was 1.29 μmol g−1 h−1 and the conversion rate of HMF was 21.6% when Pt was used as a cocatalyst. However, in the absence of Pt, the release of H2 decreased to 50 μmol g−1 h−1, and the conversion rate of HMF was only 1.5%. Therefore, Pt can be used as an effective cocatalyst in bifunctional catalytic systems. The schematic diagram is shown in Fig. 15b.
![]() | ||
| Fig. 15 (a) Comparison of H2 evolution of ZnIn2S4/NbO-8 from HMF solution and pure water, and photo-conversion of HMF in the meantime and (b) schematic illustration of photo-generated H2 release from water accompanied by the photocatalytic oxidation of HMF on ZnIn2S4/NbO-8 under simulated solar light. Reproduced with permission from ref. 113, copyright 2020, American Chemical Society. (c) Possible mechanism for photocatalytic oxidation of HMF to DFF upon ZnIn2S4-5 with abundant dual Zn and In vacancies. Reproduced with permission from ref. 114, copyright 2025, American Chemical Society. | ||
In the rational design of the catalyst for the conversion of hydrogen production coupled DFF to HMF, Zhao et al.114 also provided some ideas for the exploration of ZnIn2S4. In 2025, this team successfully created double Zn and In defects in ZnIn2S4 by a simple cetyltrimethylammonium chloride-assisted hydrothermal method. The yield (92.0%) of DFF was 3.6-fold (25.8%) higher than that of the original ZnIn2S4 when the obtained ZnIn2S4-5 was rich in cationic vacancies, and the yield of DFF was 3.6-fold higher (25.8%) than that of the original ZnIn2S4 when it was photooxidized by visible light. In particular, ZnIn2S4-5 achieved a DFF productivity of up to 1.60 mmol g−1 h−1, significantly exceeding that of the previously reported catalyst (0.075–0.95 mmol g−1 h−1) for the photooxidation of HMF in air. In addition, a mechanism is proposed whereby flower-like ZnIn2S4-5 produces photoexcited electrons and holes under visible light irradiation (Fig. 15c), in which the generation and transfer of photogenerated carriers are significantly enhanced by the double Zn and In defects. Subsequently, the O2 molecules in air are adsorbed and activated at the In site of the catalyst. At the same time, the hydroxymethyl group of HMF tends to adsorb on the Zn site of the catalyst. However, the VB energy of ZnIn2S4-5 is only 1.13 V, which is lower than the oxidation potential required for the conversion of HMF to DFF (HMF/DFF = 1.67 V vs. NHE). Therefore, the photogenerated h+ in the catalyst cannot directly oxidize HMF to DFF. In this regard, the O2 adsorbed on the surface of ZnIn2S4-5 captures the photoexcited e− and generates active ˙O2− species. Subsequently, the HMF molecule was under ˙O2− attack, forming alkoxide anions and ˙OOH intermediates. These alkoxide anions are then attacked by H+, producing alkoxide anion radicals. Finally, the α–H bond in the anionic radical was replaced by ˙OOH and is deprotonated to generate DFF and hydrogen peroxide. In the other case, the resulting ˙O2− species can be oxidized by h to 1O2 species. After that, the formed 1O2 species reacted with HMF to generate alkoxide anion radicals and ˙OOH. Eventually, through the reaction of ˙OOH with alkoxide anion radicals DFF and hydrogen peroxide are obtained.
To gain insights into the research progress of biomass oxidation coupling hydrogen production via ZnIn2S4-based photocatalysts, Table 3 summarizes recent relevant studies. It presents details such as the specific photocatalyst, synthesis method, target product, light source employed, and key performance indicators.
| Photocatalyst | Synthesis method | Product | Light source | Efficiency (mmol g−1 h−1) or selectivity | Ref. |
|---|---|---|---|---|---|
| a The E and S in the table represent the yield and selectivity of the corresponding organic compounds respectively. | |||||
| Mo2C/ZIS | Solvothermal | Furfuraldehyde | Xenon lamp (λ ≥ 420 nm 4 h) | H2: 2.80 | 115 |
| E: 11.33 | |||||
| Ni(OH)2/ZIS | in situ impregnation | Furfuraldehyde | Xenon lamp (λ ≥ 420 nm 4 h) | H2: 0.69 | 52 |
| E: 0.58 | |||||
| ReS2/ZIS | in situ growing | Furfuraldehyde | Xenon lamp (λ ≥ 420 nm 5 h) | H2: 3.09 | 116 |
| E: 2.98 | |||||
| Nb2O5/ZIS | Hydrothermal | 2, 5-diformylfuran | Xenon lamp (λ ≥ 420 nm 4 h) | H2: 1.53 | 114 |
| S: 88.30% | |||||
| O-ZIS | in situ topological transformation | 2, 5-diformylfuran | Xenon lamp (λ ≥ 420 nm 2.5 h) | H2: 1.52 | 117 |
| E: 1.62 | |||||
| CoSe2/ZIS | Hydrothermal & photodeposition | Furfuraldehyde | Xenon lamp (λ ≥ 420 nm 5 h) | H2: 46.60 | 118 |
| E: 0.85 | |||||
| CeO2/ZIS | Solvothermal | Furfuraldehyde | Xenon lamp (λ ≥ 400 nm 5 h) | H2: 1.24 | 21 |
| E: 1.34 | |||||
| ReS2/ZIS-Sv | Hydrothermal | Furfuraldehyde | Xenon lamp (λ ≥ 400 nm 2 h) | H2: 1.08 | 111 |
| E: 0.71 | |||||
First, in terms of material stability, ZnIn2S4-based matrix composites may suffer from photocorrosion under prolonged periods of light irradiation, and sulfur ions on the surface are inevitably oxidized by light and oxygen into sulfur or sulfate ions in this process. This process not only damages the integrity of the crystal structure but also leads to continuous loss of active sites. From the perspective of photocatalytic reaction mechanisms, the irrational migration of photogenerated carriers is the key factor exacerbating photo-corrosion. When photogenerated holes cannot promptly participate in surface oxidation reactions, they preferentially attack sulfur ions within the lattice. Therefore, how to fundamentally suppress the oxidation kinetics of sulfur ions by regulating carrier separation efficiency and migration pathways has become the core direction for enhancing stability. In light-corrosion-resistant design, a dual-function system integrating “carrier regulation and structural protection” should be established. By incorporating energy-matched band-structured cocatalysts, the system effectively traps photogenerated electrons while reducing sulfur ion attacks from photogenerated holes. Moreover, an ultra-thin protective layer can be formed on the surface of the catalyst by atomic layer deposition technology. This approach not only prevents direct contact between oxygen and sulfur ions but also preserves carrier transport efficiency and reactant diffusion pathways.
Second, ZnIn2S4-based photocatalysts exhibit spectral responses confined to the visible light region, with negligible utilization efficiency for near-infrared light—accounting for 43% of the solar spectrum, which severely limits their practical energy conversion performance. To achieve full-spectrum absorption and utilization, there is an urgent need to develop superior ZnIn2S4-based materials. Current research indicates that phosphorus doping and sulfur vacancy engineering can partially extend the spectral response range, though their effectiveness remains limited. From the perspective of photocatalytic material design, how to achieve efficient capture of the whole spectrum by precisely regulating the band structure or introducing the plasma resonance effect is the key to break through the bottleneck of light energy utilization. In optimizing full-spectrum response, a synergistic strategy of “defect engineering and heterojunction design” can be developed. By precisely regulating sulfur vacancy concentration, defect energy levels are introduced into the bandgap to expand the visible light response. Simultaneously, stepwise heterojunctions with narrow-bandgap semiconductors are constructed to utilize their near-infrared absorption capability, enabling photon capture across the entire spectrum. Furthermore, piezoelectric composite structures are incorporated, where mechanically induced vibrations generate built-in electric fields that enhance carrier separation, thereby indirectly improving low-light utilization efficiency.
Third, ZnIn2S4-based catalysts are still in the experimental exploration phase for hydrogen production coupling organic synthesis, particularly regarding the selective control mechanisms in complex systems (such as multi-component biomass conversion). The core advantage of photocatalytic coupling lies in utilizing photogenerated electron–hole pairs to simultaneously drive hydrogen production and organic synthesis half-reactions. However, most current research focuses solely on single reaction pathways, lacking strategies for synergistic regulation of electron transfer efficiency and reaction selectivity in coupled systems. In expanding application scenarios, the focus should be on achieving “atomic economy” and “process greening” in coupled reactions. For complex systems like biomass conversion, surface modification (such as grafting specific functional groups) can enhance catalysts' selective adsorption of target substrates. Combined with theoretical calculations (e.g., DFT simulations) to predict reaction pathways, this approach enables directed conversion from biomass to high-value chemicals. In addition, developing novel photocatalytic reactors (e.g., microchannel reactors) can not only improve mass transfer efficiency and light utilization rates, but also lay the foundation for industrial applications.
Fourth, the hydrogen production coupling organic value-added technology based on ZnIn2S4 confronts critical mechanistic challenges: the ambiguous electron distribution between hydrogen evolution sites and organic synthesis centers constrains precise control over product selectivity. The interaction mechanisms among interfacial charge migration, organic intermediates, and heterojunction energy barriers remain unclarified. And due to insufficient real-time data, the dynamic regulatory mechanisms governing hydrogen generation and organic synthesis selectivity have not been fully elucidated. To address these issues, future research should focus on the following directions. To tackle these issues, future research should concentrate on the following directions: developing integrated in situ EPR/Raman platforms to monitor electron and intermediate dynamics, thereby establishing comprehensive correlation models; employing density functional theory calculations to dissect interfacial synergistic effects, simulate electron migration barriers, and optimize cocatalyst loading and conformational configurations; and monitoring element concentration modulation strategies at active sites via in situ X-ray photoelectron spectroscopy and X-ray absorption spectroscopy techniques. Such endeavors will propel ZnIn2S4-based coupling technology from empirical optimization toward precision design, providing robust theoretical underpinnings for its practical applications. In addition, endeavors such as applying density functional theory calculations to dissect interfacial synergistic effects, simulate electron migration barriers, and optimize cocatalyst loading and conformational configurations will propel ZnIn2S4-based coupling technology from empirical optimization toward precision design, furnishing robust theoretical underpinnings for its practical applications.
| This journal is © The Royal Society of Chemistry 2025 |