A short review on research progress of ZnIn2S4-based S-scheme heterojunction: Improvement strategies

Hongrui Zhang Miaoying Cui Yongjie Lv Yongfang Rao Yu Huang

Citation:  Hongrui Zhang, Miaoying Cui, Yongjie Lv, Yongfang Rao, Yu Huang. A short review on research progress of ZnIn2S4-based S-scheme heterojunction: Improvement strategies[J]. Chinese Chemical Letters, 2025, 36(4): 110108. doi: 10.1016/j.cclet.2024.110108 shu

A short review on research progress of ZnIn2S4-based S-scheme heterojunction: Improvement strategies

English

  • The continuous growth of human demand for fossil fuels has led to energy crises and severe environmental pollution. Therefore, there is an urgent need to develop effective strategies to address these issues [1-3]. Photocatalytic technology is considered to have important application prospects in utilizing solar energy to solve energy and environmental problems [4-6]. Since Fujishiman et al. [7] discovered in 1972 that TiO2 can be used as a photocatalyst to split water to generate hydrogen gas under ultraviolet light irradiation, photocatalysis technology and semiconductor photocatalyst materials have been widely applied in various areas such as CO2 reduction [8-10], hydrogen production [11-13], and destruction of toxic and harmful pollutants [14-16].

    Researchers have developed various types of photocatalysts, such as metal oxides [17-19], graphite nitride carbon (g-C3N4) [20-22], metal sulfides [23-25], metal-organic frameworks (MOFs) [26-29], covalent organic frameworks (COFs) [30-32]. Among them, ZnIn2S4 (ZIS) has received intensive attention due to its unique layered structure, suitable bandgap, excellent visible light absorption, simple and flexible preparation methods, and good chemical stability [33]. However, ZnIn2S4, as a photocatalyst, suffers photo-generated carrier recombination and photo-corrosion, which restrains its large-scale practical application [34]. Therefore, some modification strategies have been adopted to improve the photocatalytic performance of ZnIn2S4, such as morphology regulation [35-37], element doping [38-40], vacancy engineering [41-43], and construction of heterojunctions [44-46]. Among them, constructing heterojunctions can achieve spatial separation of photo-generated carriers and effectively suppress photo-corrosion [47].

    As shown in Fig. 1, various types of ZnIn2S4-based heterojunctions have been developed, such as Schottky heterojunction, type-Ⅰ heterojunction, type-Ⅱ heterojunction, p-n heterojunction, Z-scheme heterojunction, S-scheme heterojunction, ternary and quaternary heterojunction [48]. Type-Ⅱ and Z-scheme heterojunctions have received great attention and extensive exploration. For type-Ⅱ heterojunctions, spatial separation of photo-generated carriers can be achieved, but it reduces the redox ability of photocatalysts and is not conducive to photocatalytic reactions. In addition, due to electrostatic repulsion, the transfer of electrons from the conduction band of one semiconductor to that of another semiconductor and the transfer of holes from the valence band of one semiconductor to that of another semiconductor is difficult to achieve [49]. Z-scheme heterojunctions include indirect and direct Z-scheme heterojunctions. As shown in Fig. 1F, in the case of indirect Z-scheme heterojunction, photogenerated electrons on the conduction band of semiconductors and photogenerated holes on the valence band of ZnIn2S4 are expected to migrate to the conductor for recombination. However, photogenerated electrons on the conduction band of ZnIn2S4 and holes on the valence band of semiconductor will preferentially migrate to the conductor due to a larger potential difference between conductor and the conduction band of ZnIn2S4 than that between the conduction band of semiconductor and conductor [50]. Therefore, the proposed charge transfer pathway is incorrect and the mechanism of charge transfer remains unclear. The proposed migration direction of photo-generated carriers of direct Z-scheme heterojunctions developed from indirect Z-scheme heterojunctions also remains in debate [51]. In order to tackle the controversies in the aforementioned heterojunctions, a new concept of S-scheme heterojunction was proposed based on the direct Z-scheme heterojunction [52]. Since then, there has been a surge in studies on new S-scheme heterojunctions, and many studies on ZnIn2S4-based S-scheme heterojunctions have been reported, such as g-C3N4/ZIS [53], Bi3TaO7/ZIS [54], In2O3/ZIS [55], and TpPa-1-COF/ZIS [56]. However, S-scheme heterojunctions still have some disadvantages, including poor lattice matching, limited heterojunction contact interfaces, and small Fermi-level differences [57].

    Figure 1

    Figure 1.  Illustration of (A) Schottky junction, (B) ZnIn2S4-graphene junction, (C) type-Ⅰ junction, (D) type-Ⅱ junction, (E) p-n junction, (F) indirect Z-scheme junction, (G) direct Z-scheme junction, (H) S-scheme junction, (I) ternary junction, and (J) quaternary junction. RC and OC refer to reduction cocatalyst and oxidation cocatalyst, respectively. Copied with permission [48]. Copyright 2022, EcoMat.

    Therefore, in this review, we not only introduce the design principles and characterization methods of ZnIn2S4-based S-scheme heterojunction photocatalysts, but also the improvement strategies of ZnIn2S4-based S-scheme heterojunction photocatalysts (Fig. 2). Finally, based on current research progress, we propose the challenges and future researches focuses of ZnIn2S4-based S-scheme heterojunction.

    Figure 2

    Figure 2.  The improvement methods of ZnIn2S4-based S-scheme heterojunction.

    S-scheme heterojunctions are composed of oxidation photocatalysts (OP) and reduction photocatalysts (RP) [58]. As shown in Fig. 3, RP has a smaller work function and a higher Fermi level compared to OP. When the two are in close contact, electrons in RP are spontaneously transferred to OP through the interface until the Fermi levels are the same. Therefore, at the interface, RP is positively charged due to the loss of electrons, while OP is negatively charged, thus naturally generating an internal electric field from RP to OP. Meanwhile, RP and OP exhibit upward and downward band bending at the interface, respectively. When the composite catalyst is illuminated, the electrons of RP and OP are excited from the valence band (VB) to the conduction band (CB), leaving holes in their VB, resulting in the generation of photo-generated electron-hole pairs, respectively.

    Figure 3

    Figure 3.  Charge-transfer processes in an S-scheme heterojunction. (A) before contact, (B) after contact, (C) photogenerated charge carrier transfer process in S-scheme mode. Copied with permission [51]. Copyright 2020, Elsevier lnc.

    The co-operation of internal electric field, band bending, and Coulombic force accelerates the recombination of photo-generated electrons in OP and photo-generated holes in RP, thereby suppressing the recombination of photo-generated electron-hole pairs in RP and OP, and retaining the powerful photo-generated electrons in the CB of RP and holes in the VB of OP [51]. Jiao et al. [59] selectively synthesized different types of g-C3N4 homojunctions, including CN/4ACN (type-Ⅱ), CN/8PCN (S-scheme) and 4ACN/8PCN (S-scheme), by adjusting the doping amount of Ag and P. Due to the high redox potential of S-scheme homojunctions, CN/8PCN and 4ACN/8PCN exhibited HER and degradation of tetracycline (TC). In addition, the hydrogen production rate of 4ACN/8CN was 2.1 times that of CN/4ACN. Therefore, the S-scheme heterojunction not only achieves charge separation but also exhibits the higher redox ability than type-Ⅱ heterojunction, greatly improving photocatalytic performance.

    As an n-type reduction photocatalyst, ZnIn2S4 needs to combine with another lower Fermi level oxidation photocatalyst to construct an S-scheme heterojunction in order to improve the redox ability of the composite photocatalyst [60]. In addition, ZnIn2S4 with different morphologies from 0D to 3D can be synthesized; the unique S-Zn-S-In-S-In-S layered structure of ZnIn2S4 makes its stacking structure easy to peel off and combine with another oxidation photocatalyst to construct heterojunctions [61]. Therefore, the design of ZnIn2S4-based S-scheme heterojunction photocatalysts has attracted growing interests.

    2.2.1   In situ irradiated X-ray photoelectron spectroscopy

    The change in elemental binding energy indicates the change in electron density. The decrease in electron density after losing electrons leads to an increase in binding energy, while the increase in electron density after obtaining electrons leads to a decrease in binding energy [62]. Therefore, when RP and OP are in close contact, if an S-scheme heterojunction is successfully constructed, due to the difference in their work functions, the electrons of RP will transfer to OP until the Fermi level is the same, manifested as the binding energies of RP shifting toward higher energy levels, while the binding energies of OP moving toward lower energy levels. When illuminated, the photogenerated electrons in the OP conduction band will migrate to RP and recombine with the photogenerated holes in the RP valence band, manifested as the binding energies of RP shifting toward lower energy levels, while the binding energies of OP shifting toward higher energy levels [63]. The change in the binding energy of in-situ XPS suggests the transfer direction of charge carriers in heterojunctions, providing direct evidence for the successful construction of S-scheme heterojunctions.

    2.2.2   Atomic force microscope (AFM)

    Kelvin probe force microscopy (KPFM) is a derivative of AFM, which can scan the morphology of AFM while detecting its surface potential and can be used to evaluate the charge transfer process [13]. Yu et al. [64] constructed inorganic/organic heterojunctions by in-situ growth of CdS nanocrystals on the surface of pyrene-alt-triphenylamine (PT), and demonstrated the charge transfer mechanism of S-scheme heterojunctions between CdS/PT through KPFM, with PT serving as a reducing photocatalyst and CdS as an oxidizing photocatalyst. As shown in Fig. 4, when CdS and PT come into close contact, electrons will transfer from PT to CdS. Therefore, under dark conditions, the surface potential difference between PT (point A) and CdS (point B) is about 100 mV, indicating the formation of an internal electric field from point A to point B. When illuminated by light, the surface potential of point A decreases significantly, while the surface potential of point B increases, indicating the transfer of photo-generated electrons from CB of CdS to VB of PT. Therefore, KPFM characterization technology can also be used to characterize the charge transfer pathway in S-scheme heterojunctions.

    Figure 4

    Figure 4.  (A) Atomic force microscopy image of CP composite, (B, C) corresponding surface potential distribution of CP in darkness (B) and under light irradiation (C). (D) The line-scanning surface potential from point A to B. (E) The schematic illustration of photoirradiation KPFM. Copied with permission [64]. Copyright 2021, Wiley-VCH GmbH.
    2.2.3   Electron paramagnetic resonance (EPR)

    Since the successful construction of S-scheme heterojunctions leads to the recombination of photo-generated carriers with weak redox ability, and the strong redox ability of composite photocatalysts is retained, the formation of S-scheme heterojunctions can be confirmed by EPR detection of free radicals generated during the photocatalytic process, such as hydroxyl and superoxide radicals [65]. Yu et al. [66] designed a heterojunction with a core-shell structure of TiO2@ZnIn2S4 for efficient photocatalytic reduction of CO2, and the type of heterojunction was determined through EPR test. As shown in Fig. 5A, based on the band structures of TiO2 and ZnIn2S4, the electrons on the CB of ZnIn2S4 can reduce O2 to O2, but the holes on the VB of ZnIn2S4 cannot oxidize H2O to OH; while the electrons on the CB of TiO2 cannot reduce O2 to O2, but the holes on the VB of TiO2 can oxidize H2O to OH.

    Figure 5

    Figure 5.  (A) Band structures of samples. Electron paramagnetic resonance (EPR) spectra of (B) DMPO-OH adducts (in aqueous solution) and (C) DMPO-O2 adducts (in methanol solution) over TiO2, ZnIn2S4, and TiO2@ZnIn2S4 under illumination for 2 min. Schematic diagrams of (D) conventional type-Ⅱ and (E) S-scheme heterojunction. Reproduced with permission [66]. Copyright 2021, Wiley-VCH GmbH.

    Therefore, as shown in Figs. 5B and C, the DMPO-OH signal and the extremely weak DMPO-O2 signal over TiO2, as well as the DMPO-O2 signal over ZnIn2S4, can be observed in the EPR spectra, while the DMPO-OH signal over ZnIn2S4 cannot be detected. In the case of TiO2@ZnIn2S4, the intensity of DMPO-OH and DMPO-O2 signals is stronger than that over the sole-TiO2 and sole-ZnIn2S4, respectively. If TiO2@ZnIn2S4 is a type-Ⅱ heterojunction as shown in Fig. 5D, both OH and O2 cannot be generated. Therefore, as shown in Fig. 5E, TiO2@ZnIn2S4 is an S-scheme heterojunction, retaining strong oxidation and reduction capability of both, resulting in the production of both OH and O2.

    Efficient photocatalytic performance can be achieved by reasonably designing S-scheme heterojunctions with fast interface charge transfer and high redox potential [67]. However, S-scheme heterojunctions are usually composed of two n-type semiconductors, and the Fermi energy level difference between the two components is small, resulting in weak electric field driving force for interfacial charge transfer, low charge separation efficiency, and thus, poor photocatalytic performance [68]. Therefore, some modification strategies, such as morphological control, defect engineering, element doping, have been applied to provide S-scheme heterojunctions with more active sites and band arrangements with larger interface potential differences, thereby promoting charge separation and improving activity. The improvement strategies of ZnIn2S4-based S-scheme heterojunction are summarized in Table 1.

    Table 1

    Table 1.  Summary table of the improvement methods of ZnIn2S4-based S-scheme heterojunction.
    DownLoad: CSV

    By reasonable morphological design, photocatalytic activity of S-scheme heterojunction can be greatly improved [69]. Dai et al. [70] synthesized the S-scheme heterojunction CdLa2S4/ZnIn2S4 by loading fishbone-like CdLa2S4 on micrometer flower-shaped ZIS based on different work functions. The unique structure provided a large contact area between the catalyst and the reaction solution. Through careful design of the morphology and the construction of S-scheme heterojunction, the degradation efficiency of sulfamethoxazole was as high as 98.9%.

    However, the 2D sheet-like structure of ZnIn2S4 often undergoes excessive self-assembly to produce flower-shaped microspheres, which greatly reduces the active sites and leads to photo-generated carrier recombination [71]. Therefore, constructing heterojunctions between catalysts with appropriate morphology and ZIS nanosheets can prevent aggregation and improve photocatalytic activity [72]. Zhao et al. [73] constructed the S-scheme heterojunction ZIS/WO3 by in-situ growth of 2D ZIS nanosheets on 1D WO3 nanorods. The 1D nanorod structure can avoid the aggregation of ZIS sheet-like structures and has a large specific surface area and a short charge transfer path. The vertical distribution of 2D nanosheets on the surface of 1D nanorods can improve the surface activity and light absorption of photocatalysts. Thanks to the S-scheme heterojunction and exquisite morphology design, the ZIS/WO3 heterojunction exhibited excellent photocatalytic activity, with a hydrogen production rate of 300 µmol g−1 h−1.

    Compared to 1D/2D composite, 2D/2D composite not only prevents the aggregation of ZIS nanosheets, but also the OP substrate of 2D can provide a larger specific surface area and tighter surface contact, thereby providing more adsorption and active sites and rich charge transfer channels for the construction of S-scheme heterojunctions [74]. Wang et al. [75] prepared mossy tile-like morphology S-scheme heterojunction ZIS/CMS by in-situ growth of ZIS nanosheets on Cu2MoS4 plates. The unique structure inhibited the aggregation of ZIS nanosheets, provided more active sites, and led to the formation of a tight interface S-scheme heterojunction between ZIS and CMS. The photocatalytic hydrogen production rate of ZIS/CMS was as high as 1298 µmol g−1 h−1 and had excellent stability. In addition, due to the unique layered crystal structure of Bi4Ti3O12, Zhou et al. [76] prepared an S-scheme heterojunction 2D/2D ZnIn2S4/Bi4Ti3O12 using molten salt method and low-temperature solvothermal method. Benefiting from the abundant active sites, tight contact interfaces, and effective separation of photo-generated carriers, ZIS/BTO exhibited good tetracycline removal efficiency, with a rate constant of 0.023 min−1.

    Compared to other morphology engineering, hollow structures can offer nanomaterials some unique advantages including large specific surface area, a great number of active sites, and improved light utilization through multiple reflections and scattering of light [77]. Li et al. [78] prepared an S-scheme heterojunction ZIS/HCNT with core-shell structure by in-situ growth of ZIS nanosheets on g-C3N4 hollow nanotubes. Compared to 1D nanorods, 1D nanotubes had higher light utilization efficiency and stronger charge transfer ability due to multiple reflection effects and thin tube walls. In situ growth promoted the formation of a uniform and tight interface between the two, and simultaneously utilized the advantages of hollow structure and S-scheme heterojunction, exhibiting a high CO yield of 883 µmol g−1 h−1.

    For S-scheme heterojunctions, increasing the difference in work function between RP and OP can also enhance the strength of the internal electric field, thereby promoting the transfer of interfacial charges in heterojunctions and the separation of photo-generated carriers in photocatalysts [79]. The introduction of vacancies can regulate the work function of catalysts, thus improving S-scheme heterojunctions through defect engineering [80]. Liu et al. [81] first prepared ZIF-8/ZIS with atomic level tight interfaces through the ZIS self-sacrificing reaction mechanism and then controlled the in-situ sulfurization time to convert ZIF-8/ZIS into an S-scheme heterojunction Sv-ZnS/ZIS with controllable sulfur vacancy concentration. The introduction of sulfur vacancies changed the work function of ZnS, increasing the difference in work function between ZIS and ZnS from 1.41 eV to 1.53 eV. Sv-ZnS/ZIS with enhanced internal electric field strength and tight interface exhibited excellent photocatalytic hydrogen production performance; the hydrogen production rate of Sv-ZnS/ZIS was 2912.3 ± 185.9 µmol g−1 h−1, which was much higher than the hydrogen production rate of other ZnS/ZIS photocatalyst by 103 µmol g−1 h−1 [82].

    The bridging of interface chemical bonds can provide a charge transmission channel, which is conducive to faster electron transfer between interfaces and further expands the advantages provided by defect engineering [94]. Sun et al. [83] combined g-C3N4 with ZIS with sulfur vacancies to construct an S-scheme heterojunction Sv-ZIS/CN with efficient degradation of tetracycline. Within 90 min, the tetracycline degradation rate of Sv-ZIS/CN reached 96.36%, much higher than the 64.62% tetracycline degradation rate of ZIS-S/CN prepared by mechanical stirring. The excellent activity was attributed to the introduction of sulfur vacancies, the formation of interface S-C bonds, and the internal electric field, which jointly promoted charge transfer in S-scheme heterojunctions and significantly improved the separation efficiency of photo-generated carriers.

    The combination of S-scheme heterojunction and element doping can synergistically improve light absorption and promote the separation of photo-generated carriers, thereby enhancing photocatalytic performance. Su et al. [85] utilized the thermal solubility of MoO3 to obtain uniform in-situ Mo doped ZIS wrapped MoO3 S-scheme heterojunction. In situ Mo doping not only improved light absorption, but also led to the formation of Mo-S bonds, reducing the ΔGH*, leading to excellent photocatalytic hydrogen production performance (5.5 mmol g−1 h−1), which was much better than that of MoO3@ZIS (2.4 mmol g−1 h−1).

    Element doping can achieve a larger Fermi energy level difference between heterojunctions by adjusting the work function, resulting in a stronger internal electric field and higher interfacial charge transfer efficiency [95]. In addition, elemental doping can also change the hydrophilicity of the catalyst surface, promoting the adsorption and activation of H2O molecules [96]. Zhu et al. [84] constructed an S-scheme heterojunction by in-situ growth of Mo-modified ZnIn2S4 nanosheets on rod-shaped NiTiO3 using the hydrothermal method (Mo-ZIS@NTO). Mo doping significantly increased the Fermi level of ZIS, resulting in an increase in the Fermi level difference between ZIS and NTO from 1.0 eV to 1.43 eV. The enhanced internal electric field as a powerful driving force can promote the separation and transmission of photo-generated charge carriers. The increased hydrophilicity of Mo-ZIS helped to adsorb more water molecules for photocatalytic reduction of hydrogen production. Through the combined effect of S-scheme heterojunction and Mo doping, Mo-ZIS@NTO exhibited excellent performance, with a hydrogen production rate of 14.06 mmol g−1 h−1 under visible light, which was 2.2 times higher than ZIS@NTO.

    For S-scheme heterojunction structures, the transfer and spatial separation of photo-generated carriers in RP and OP occur between interfaces. Different crystal planes of the same crystal phase will have different energy band levels due to differences in their surface electronic structures [97]. Therefore, the interface contact between ZIS and other semiconductors with different facet exposure can lead to heterojunctions with different band structures, thereby affecting the migration of photo-generated carriers between interfaces [98]. Xi et al. [86] used crystal plane engineering to adjust the arrangement of energy bands, regulate charge transfer and separation between interfaces, and significantly improved the photocatalytic activity of S-scheme heterojunctions for hydrogen production. ZIS/BOB-(010) and ZIS/BOB-(001) S-scheme heterojunctions were prepared by growing ZIS nanosheets on BiOBr nanoparticles mainly composed of (010) and (001) facet, respectively. The Fermi energy level difference (0.70 eV) between BOB-(001) and ZIS was larger than that (0.49 eV) between BOB-(010) and ZIS. Due to the larger Fermi energy level difference between BOB-(001) and ZIS, the interface exhibited more significant band bending and stronger internal electric field, accelerating the transfer efficiency of photo-generated carriers through the BOB-(001)/ZIS interface, thereby promoting the recombination of useless photo generated carriers with weak redox ability and the spatial separation of useful photogenerated electron-hole pairs with strong redox ability. Therefore, BOB-(001)/ZIS exhibited excellent photocatalytic hydrogen production efficiency, up to 17.0 mmol g−1 h−1, which was 4.0 and 2.4 times higher than that over ZIS and BOB-(010)/ZIS, respectively.

    As a temperature sensitive photocatalyst, the increase in temperature leads to an increase in its activity of ZnIn2S4 [88]. Therefore, coupling ZnIn2S4 nanosheets with photothermal materials to construct an S-scheme heterojunction can achieve higher photocatalytic activity, which is beneficial for the development of photothermal-assisted photocatalytic systems [91]. Chen et al. [87] constructed an S-scheme heterojunction by in-situ growth of ZIS nanosheets on FeS2 hollow spheres, a photothermal material. The photothermal effect of FeS2 increased the temperature of the reaction system and accelerated charge transfer; the surface ZIS coating could effectively prevent heat loss. Under simulated solar radiation, the hydrogen production rate over FS2@ZIS was as high as 5.05 mmol g−1 h−1, which was 47.9 and 53.7 times that over the original ZIS and FeS2, respectively.

    In recent years, the localized surface plasmon resonance (LSPR) effect has also been found in some non-stoichiometric compounds, enabling them to possess the ability of photothermal synergistic catalytic reactions [90]. Xiong et al. [89] successfully synthesized the S-scheme heterojunction MoO3-x@ZIS by in-situ growth of ZIS nanosheets on 1D needle-like MoO3-x. Due to the large number of oxygen vacancies in MoO3-x, Mo6+ was reduced to Mo5+, resulting in an increase in free electron concentration and the LSPR effect. Under the synergistic effect of S-scheme heterojunction and photothermal conversion, the reaction efficiency and CH4 selectivity of CO2 reduction were improved. Under full spectrum irradiation, the yields of CO and CH4 over MoO3-x@ZIS reached 4.65 and 28.3 µmol g−1 h−1, respectively, and CH4 selectivity was as high as 85.89%, with CH4 yield, 19.4 and 11.7 times higher than that over MoO3-x and ZIS.

    Compared to binary S-scheme heterojunctions with single-channel charge transfer, charge transfer of multicomponent heterojunctions with multi-carrier transport paths is more efficient [99]. Bai et al. [100] anchored the S-scheme heterojunction ZIS-NiSe2 onto Ti3C2 MXene using a two-step solvothermal method. Due to the higher Fermi level of ZIS compared to NiSe2 and Ti3C2 MXene, the charge of ZIS will be transferred to NiSe2 and Ti3C2 MXene. When the Fermi levels of the three were balanced, an internal electric field was generated between the interface of ZIS and NiSe2, and a Schottky barrier was formed between the interface of ZIS and Ti3C2 MXene. When illuminated, the photo-generated electrons in Ti3C2 CB will recombine with the photo-generated holes in ZIS VB in the internal electric field, while the photo-generated electrons on ZIS CB will migrate along the Schottky junction to Ti3C2 MXene, thereby suppressing the photo-generated electron reflux in ZIS. Under the synergistic effect of Schottky junction and S-scheme heterojunction, the dual charge transfer pathways in ZIS can achieve rapid photo-generated carrier separation; Ti3C2-ZIS-NiSe2 exhibited a hydrogen production rate of 23.51 mmol g−1 h−1 under visible light, which was 23.51, 8.31, and 3.54 times higher than that over ZIS, Ti3C2-ZIS, and ZIS-NiSe2, respectively.

    Xia et al. [92] promoted the charge transfer of 2D/2D S-scheme heterojunctions without close contact region by inserting 0D CNQDs between 2D/2D TCN (tubular g-C3N4)/ZIS S-scheme heterojunctions to construct Schottky junctions. The synergistic effect of this Schottky junction and S-scheme heterojunction greatly improved the transfer and separation of charges between the 2D/2D TCN/ZIS interfaces, generating abundant OH, O2 and photo-generated holes to remove carbon atoms in C14H30 one by one, ultimately efficiently degrading n-tetradecane into CO2 and H2O.

    Most studies on S-scheme photocatalysts have focused on binary composite materials. However, there is still considerable potential for improvement in charge transfer, quantum efficiency, and light capture ability of the binary S-scheme photocatalyst [101]. The dual S-scheme heterojunction can further optimize the photogenerated charge transfer pathway, and its excellent photocatalytic activity can be attributed to the synergistic effect between the three components [102]. Liu et al. [93] used the isoelectric point assisted calcination method to fix g-C3N4, TiO2, and ZIS on the graphene aerogel carrier and constructed a dual S-scheme heterojunction g-C3N4/TiO2/ZIS (CTZA) which efficiently degraded methyl orange (MO) under visible light irradiation. Due to the lower Fermi level of TiO2 compared to g-C3N4 and ZIS, when the three came into contact, the charges on g-C3N4 and ZIS will transfer to TiO2 until the Fermi level was balanced, and a dual internal electric field directed by g-C3N4 and ZIS towards TiO2 was formed. Under illumination, the photo-generated electrons in the CB of TiO2 recombined with the photo-generated holes in g-C3N4 and ZIS, thereby retaining the holes with strong oxidation ability in the VB of TiO2 and the electrons with strong reduction ability in g-C3N4 and ZIS. Graphene with excellent conductivity can further promote the transfer and separation of photo-generated charge carriers, resulting in a degradation rate of 97.5% for MO over CTZA within 30 min, much higher than that over ZIS, TiO2 and g-C3N4.

    ZnIn2S4, as a reducing ternary metal sulfide photocatalyst, has received widespread attention due to its narrow bandgap and visible-light-responsive properties. Unfortunately, the rapid recombination of photo-generated carriers and photo-corrosion of ZnIn2S4 curbs its practical application. Although constructing S-scheme heterojunctions by combining with other semiconductor materials can improve the photocatalytic activity of ZnIn2S4, developing effective strategies which can further improve the photocatalytic performance of ZnIn2S4-based S-scheme heterojunctions still remain a challenge. This article introduces the concept and design principles of ZnIn2S4-based S-scheme heterojunction, and summarizes its current improvement strategies for S-scheme heterojunction, including morphology control, defect engineering, element doping, facet engineering, photothermal assisted catalysis, and ternary heterojunction. Although the studies on the improvement of ZnIn2S4-based S-scheme heterojunctions have made significant progress, a few research areas can be focused in the future.

    Firstly, to identify the direction of charge transfer in S-scheme heterojunctions more accurately, more advanced in-situ characterization techniques should be applied to demonstrate the charge transfer process besides in-situ XPS and AFM. For instance, time-resolved photoemission electron microscopy (TR-PEEM) can be applied to visually showcase the transfer of electrons and holes of ZnIn2S4-based S-scheme heterojunctions.

    Secondly, combining experiments with theoretical simulation calculations helps understand the dynamics of charge carrier transfer and identify the active sites of OP and RP, which benefits the rational design of photocatalysts for the selective reaction. For example, for photocatalytic reduction of CO2 using ZnIn2S4-based S-scheme heterojunctions, active sites favorable for the generation of reaction intermediates can be predicted through theoretical calculations, and the electron enrichment ability of reducing sites can be increased by improving the S-scheme heterojunction. Therefore, rational design of S-scheme heterojunctions can further enhance product selectivity and generate more valuable hydrocarbon fuels.

    Thirdly, previous studies focused on the improvement of ZnIn2S4-based S-scheme heterojunctions through anionic vacancies and metal doping. Cation vacancies, non-metal doping and double doping can also be applied to improve the performance of S-scheme heterojunctions in the future.

    Finally, extending the photo-response of ZnIn2S4-based S-scheme heterojunction to near-infrared range through narrowing the bandgap and combining ZnIn2S4-based S-scheme heterojunction with up-conversion materials, photothermal-conversion materials or photosensitive materials can also be a research focus since near-infrared light accounts for 52% of solar light.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Hongrui Zhang: Writing – original draft. Miaoying Cui: Investigation. Yongjie Lv: Investigation. Yongfang Rao: Writing – review & editing. Yu Huang: Resources.

    We are grateful to the Strategic Priority Research Program of the Chinese Academy of Sciences, China (Nos. XDA23010300 and XDA23010000) and the National Natural Science Foundation of China (Nos. 51878644 and 41573138).


    1. [1]

      Y. Kumar, R. Kumar, P. Raizada, et al., J. Mater. Sci. Technol. 87 (2021) 234–257. doi: 10.1016/j.jmst.2021.01.051

    2. [2]

      W.J. Liu, H. Jiang, H.Q. Yu, Energy Environ. Sci. 12 (2019) 1751–1779. doi: 10.1039/c9ee00206e

    3. [3]

      H.J. Huang, M.M. Yan, C.Z. Yang, et al., Adv. Mater. 31 (2019) 1903415. doi: 10.1002/adma.201903415

    4. [4]

      Q. Guo, C.Y. Zhou, Z.B. Ma, et al., Adv. Mater. 31 (2019) 1901997. doi: 10.1002/adma.201901997

    5. [5]

      F. Chen, H.W. Huang, L. Guo, et al., Angew. Chem. Int. Ed. 58 (2019) 10061–10073. doi: 10.1002/anie.201901361

    6. [6]

      Y. Cai, F.X. Luo, Y.J. Guo, et al., Molecules 28 (2023) 2142. doi: 10.3390/molecules28052142

    7. [7]

      A. Fujishima, K. Honda, Nature 238 (1972) 37–38. doi: 10.1038/238037a0

    8. [8]

      X.J. Shi, Y. Huang, Y.N. Bo, et al., Angew. Chem. Int. Ed. 61 (2022) e202203063. doi: 10.1002/anie.202203063

    9. [9]

      H. Li, C.K. Cheng, Z.J. Yang, et al., Nat. Commun. 13 (2022) 6466. doi: 10.1038/s41467-022-34263-z

    10. [10]

      Y.N. Zhang, M.Y. Xu, W.Q. Zhou, et al., J. Colloid Interf. Sci. 650 (2023) 1762–1772. doi: 10.1016/j.jcis.2023.07.120

    11. [11]

      X.J. Chen, J. Wang, Y.Q. Chai, et al., Adv. Mater. 33 (2021) 2007479. doi: 10.1002/adma.202007479

    12. [12]

      M.L. Xu, M. Lu, G.Y. Qin, et al., Angew. Chem. Int. Ed. 61 (2022) e202210700. doi: 10.1002/anie.202210700

    13. [13]

      B.Q. Xia, B.W. He, J.J. Zhang, et al., Adv. Energy Mater. 12 (2022) 2201449. doi: 10.1002/aenm.202201449

    14. [14]

      X.Y. Chen, X. Peng, L.B. Jiang, et al., Chem. Eng. J. 427 (2022) 130945. doi: 10.1016/j.cej.2021.130945

    15. [15]

      M. Dai, Z.L. He, W.R. Cao, et al., Sep. Purif. Technol. 309 (2023) 123004. doi: 10.1016/j.seppur.2022.123004

    16. [16]

      Y. Wang, X. Wu, J.Q. Liu, et al., J. Environ. Chem. Eng. 10 (2022) 107091. doi: 10.1016/j.jece.2021.107091

    17. [17]

      Y. Yuan, R.T. Guo, L.F. Hong, et al., Mater. Today Energy 21 (2021) 100829. doi: 10.1016/j.mtener.2021.100829

    18. [18]

      X. Li, H.P. Jiang, C.C. Ma, et al., Appl. Catal. B: Environ. 283 (2021) 119638. doi: 10.1016/j.apcatb.2020.119638

    19. [19]

      E.S. Ghalehsefid, Z.G. Jahani, A. Aliabadi, et al., J. Environ. Chem. Eng. 11 (2023) 110160. doi: 10.1016/j.jece.2023.110160

    20. [20]

      B.C. Zhu, B. Cheng, J.J. Fan, et al., Small Struct. 2 (2021) 2100086. doi: 10.1002/sstr.202100086

    21. [21]

      Y.F. Huang, D. Wei, Z.Y. Li, et al., J. Energy Chem. 83 (2023) 423–432. doi: 10.1016/j.jechem.2023.04.008

    22. [22]

      X.H. Wu, X.F. Wang, F.Z. Wang, et al., Appl. Catal. B: Environ. 247 (2019) 70–77. doi: 10.1016/j.apcatb.2019.01.088

    23. [23]

      G.X. Zhang, Z.J. Guan, J.J. Yang, et al., Sol. RRL 6 (2022) 2200587. doi: 10.1002/solr.202200587

    24. [24]

      J. Jiang, G.H. Wang, Y.C. Shao, et al., Chin. J. Catal. 43 (2022) 329–338. doi: 10.1016/S1872-2067(21)63889-5

    25. [25]

      K. Wang, X.L. Shao, Q. Cheng, et al., Sol. RRL 6 (2022) 2200736. doi: 10.1002/solr.202200736

    26. [26]

      N.Y. Huang, J.Q. Shen, X.W. Zhang, et al., J. Am. Chem. Soc. 144 (2022) 8676–8682. doi: 10.1021/jacs.2c01640

    27. [27]

      K. Sun, Y.Y. Qian, H.L. Jiang, Angew. Chem. Int. Ed. 62 (2023) e202217565. doi: 10.1002/anie.202217565

    28. [28]

      Y.P. Zhang, J.X. Xu, J. Zhou, et al., Chin. J. Catal. 43 (2022) 971–1000. doi: 10.1016/S1872-2067(21)63934-7

    29. [29]

      R.N. Ali, W.A. Qureshi, H. Naz, et al., New J. Chem. 47 (2023) 15534–15542. doi: 10.1039/d3nj02143b

    30. [30]

      R.F. Chen, Y. Wang, Y. Ma, et al., Nat. Commun. 12 (2021) 1354. doi: 10.1038/s41467-021-21527-3

    31. [31]

      J.H. Ding, X.Y. Guan, J. Lv, et al., J. Am. Chem. Soc. 145 (2023) 3248–3254. doi: 10.1021/jacs.2c13817

    32. [32]

      L. Sun, L.L. Li, J. Yang, et al., Chin. J. Catal. 43 (2022) 350–358. doi: 10.1016/S1872-2067(21)63869-X

    33. [33]

      Y.J. Ren, J.J. Foo, D.Q. Zeng, et al., Small Struct. 3 (2022) 2200017. doi: 10.1002/sstr.202200017

    34. [34]

      T.X. Zhang, T. Wang, F.L. Meng, et al., J. Mater. Chem. C 10 (2022) 5400–5424. doi: 10.1039/d2tc00432a

    35. [35]

      X.L. Gou, F.Y. Cheng, Y.H. Shi, et al., J. Am. Chem. Soc. 128 (2006) 7222–7229. doi: 10.1021/ja0580845

    36. [36]

      M.S. Yu, X.Y. Lv, A.M. Idris, et al., J. Colloid Interf. Sci. 612 (2022) 782–791. doi: 10.1016/j.jcis.2021.12.197

    37. [37]

      S.P. Ding, X.F. Liu, Y.Q. Shi, et al., ACS Appl. Mater. Interfaces 10 (2018) 17911–17922. doi: 10.1021/acsami.8b02955

    38. [38]

      B. Pan, Y. Wu, B. Rhimi, et al., J. Energy Chem. 57 (2021) 1–9. doi: 10.1155/2021/8132569

    39. [39]

      D.X. Zhou, X.D. Xue, X. Wang, et al., Appl. Catal. B: Environ. 310 (2022) 121337. doi: 10.1016/j.apcatb.2022.121337

    40. [40]

      S.Q. Zhang, Z.F. Zhang, Y.M. Si, et al., ACS Nano 15 (2021) 15238–15248. doi: 10.1021/acsnano.1c05834

    41. [41]

      Y.Q. He, C.L. Chen, Y.X. Liu, et al., Nano Lett. 22 (2022) 4970–4978. doi: 10.1021/acs.nanolett.2c01666

    42. [42]

      Y.Q. He, H. Rao, K.P. Song, et al., Adv. Funct. Mater. 29 (2019) 1905153. doi: 10.1002/adfm.201905153

    43. [43]

      J.W. Pan, G.X. Zhang, Z.J. Guan, et al., J. Energy Chem. 58 (2021) 408–414. doi: 10.1016/j.jechem.2020.10.030

    44. [44]

      Q.L. Huang, J.Q. Hu, Y.F. Hu, et al., Environ. Sci. Nano 9 (2022) 4433–4444. doi: 10.1039/d2en00781a

    45. [45]

      J.M. Li, C.C. Wu, J. Li, et al., Chin. J. Catal. 43 (2022) 339–349. doi: 10.1016/S1872-2067(21)63875-5

    46. [46]

      S. Cao, J.G. Yu, S. Wageh, et al., J. Mater. Chem. A 10 (2022) 17174–17184. doi: 10.1039/d2ta05181h

    47. [47]

      Q.H. Zhu, Q. Xu, M.M. Du, et al., Adv. Mater. 34 (2022) 2202929. doi: 10.1002/adma.202202929

    48. [48]

      V.B.Y. Oh, S.F. Ng, W.J. Ong, Ecomat 4 (2022) e12204. doi: 10.1002/eom2.12204

    49. [49]

      L.Y. Zhang, J.J. Zhang, H.G. Yu, et al., Adv. Mater. 34 (2022) 2107668. doi: 10.1002/adma.202107668

    50. [50]

      S. Wageh, A.A. Al-Ghamdi, R. Jafer, et al., Chin. J. Catal. 42 (2021) 667–669. doi: 10.1016/S1872-2067(20)63705-6

    51. [51]

      Q.L. Xu, L.Y. Zhang, B. Cheng, et al., Chem 6 (2020) 1543–1559. doi: 10.1016/j.chempr.2020.06.010

    52. [52]

      J.W. Fu, Q.L. Xu, J.X. Low, et al., Appl. Catal. B: Environ. 243 (2019) 556–565. doi: 10.1016/j.apcatb.2018.11.011

    53. [53]

      Y.N. Zhang, M. Gao, S.T. Chen, et al., Acta Phys. Chim. Sin. 39 (2023) 2211051. doi: 10.3866/pku.whxb202211051

    54. [54]

      K. Wang, X.L. Shao, K.J. Zhang, et al., Appl. Surf. Sci. 596 (2022) 153444. doi: 10.1016/j.apsusc.2022.153444

    55. [55]

      L. Zhao, B.X. Yang, G.X. Zhuang, et al., Small 18 (2022) 2201668. doi: 10.1002/smll.202201668

    56. [56]

      P.Y. Dong, T. Cheng, J.L. Zhang, et al., ACS Appl. Energy Mater. 6 (2023) 1103–1115. doi: 10.1021/acsaem.2c03806

    57. [57]

      A. Kumar, A. Khosla, S.K. Sharma, et al., Fuel 333 (2023) 126267. doi: 10.1016/j.fuel.2022.126267

    58. [58]

      Q.L. Xu, S. Wageh, A.A. Al-Ghamdi, et al., J. Mater. Sci. Technol. 124 (2022) 171–173. doi: 10.1016/j.jmst.2022.02.016

    59. [59]

      J.J. Jiao, Z.X. Wang, J. Wang, et al., Appl. Surf. Sci. 639 (2023) 158260. doi: 10.1016/j.apsusc.2023.158260

    60. [60]

      Z.F. Yang, L.L. Wang, M. Fang, et al., Sep. Purif. Technol. 305 (2023) 122509. doi: 10.1016/j.seppur.2022.122509

    61. [61]

      J. Wang, S.J. Sun, R. Zhou, et al., J. Mater. Sci. Technol. 78 (2021) 1–19. doi: 10.1016/j.jmst.2020.09.045

    62. [62]

      Z.Y. Xie, L.J. Xie, F.F. Qi, et al., J. Colloid Interf. Sci. 650 (2023) 784–797. doi: 10.1016/j.jcis.2023.07.032

    63. [63]

      J.J. Zhang, L.Y. Zhang, W. Wang, et al., J. Phys. Chem. Lett. 13 (2022) 8462–8469. doi: 10.1021/acs.jpclett.2c02125

    64. [64]

      C. Cheng, B.W. He, J.J. Fan, et al., Adv. Mater. 33 (2021) 2100317. doi: 10.1002/adma.202100317

    65. [65]

      X.N. Wang, M. Sayed, O. Ruzimuradov, et al., Appl. Mater. Today 29 (2022) 101609. doi: 10.1016/j.apmt.2022.101609

    66. [66]

      L.B. Wang, B. Cheng, L.Y. Zhang, et al., Small 17 (2021) 2103447. doi: 10.1002/smll.202103447

    67. [67]

      X.L. Su, S.K. Wang, J.C. Liu, et al., Chemosphere 340 (2023) 139777. doi: 10.1016/j.chemosphere.2023.139777

    68. [68]

      J.Q. Liu, J. Wan, L. Liu, et al., Chem. Eng. J. 430 (2022) 133125. doi: 10.1016/j.cej.2021.133125

    69. [69]

      S. Wang, X. Du, C.H. Yao, et al., Nano Res. 16 (2023) 2152–2162. doi: 10.1007/s12274-022-4960-8

    70. [70]

      M. Dai, H.J. Yu, W.H. Chen, et al., Chem. Eng. J. 470 (2023) 144240. doi: 10.1016/j.cej.2023.144240

    71. [71]

      J.J. Peng, J.Y. Yang, B. Chen, et al., Biosens. Bioelectron. 175 (2021) 112873. doi: 10.1016/j.bios.2020.112873

    72. [72]

      H.T. Fan, Z. Wu, K.C. Liu, et al., Chem. Eng. J. 433 (2022) 134474. doi: 10.1016/j.cej.2021.134474

    73. [73]

      M.Y. Zhao, S. Liu, D.M. Chen, et al., Chin. J. Catal. 43 (2022) 2615–2624. doi: 10.1016/S1872-2067(22)64134-2

    74. [74]

      Z.T. He, C. Qian, D.M. Chen, et al., J. Colloid Interf. Sci. 651 (2023) 138–148. doi: 10.1016/j.jcis.2023.07.179

    75. [75]

      S.K. Wang, D.F. Zhang, P. Su, et al., J. Colloid Interf. Sci. 650 (2023) 825–835. doi: 10.1016/j.jcis.2023.07.052

    76. [76]

      Q. Zhou, L.H. Zhang, L.F. Zhang, et al., J. Hazard. Mater. 438 (2022) 129438. doi: 10.1016/j.jhazmat.2022.129438

    77. [77]

      Y.C. Guo, J.M. Sun, Y. Tang, et al., Energy Environ. Sci. 16 (2023) 3462–3473. doi: 10.1039/d3ee01522j

    78. [78]

      L.L. Li, D.K. Ma, Q.L. Xu, et al., Chem. Eng. J. 437 (2022) 135153. doi: 10.1016/j.cej.2022.135153

    79. [79]

      Y.X. Tan, Z.M. Chai, B.H. Wang, et al., ACS Catal. 11 (2021) 2492–2503. doi: 10.1021/acscatal.0c05703

    80. [80]

      Z.J. Zhang, X.S. Wang, D.B. Li, et al., Small 20 (2024) 2305566. doi: 10.1002/smll.202305566

    81. [81]

      J.Y. Liu, M. Liu, S.B. Zheng, et al., J. Colloid Interf. Sci. 635 (2023) 284–294. doi: 10.1016/j.jcis.2022.12.131

    82. [82]

      Y.X. Li, J.X. Wang, S.Q. Peng, et al., Int. J. Hydrogen Energy 35 (2010) 7116–7126. doi: 10.1016/j.ijhydene.2010.02.017

    83. [83]

      L.L. Sun, X.S. Liu, X.H. Jiang, et al., J. Mater. Chem. A 10 (2022) 25279–25294. doi: 10.1039/d2ta08337j

    84. [84]

      J.F. Zhu, Q.Y. Bi, Y.H. Tao, et al., Adv. Funct. Mater. 33 (2023) 2213131. doi: 10.1002/adfm.202213131

    85. [85]

      H. Su, H.M. Lou, Z.P. Zhao, et al., Chem. Eng. J. 430 (2022) 132770. doi: 10.1016/j.cej.2021.132770

    86. [86]

      Y.M. Xi, W.B. Chen, W.R. Dong, et al., ACS Appl. Mater. Interfaces 13 (2021) 39491–39500. doi: 10.1021/acsami.1c11233

    87. [87]

      K.Y. Chen, Y.X. Shi, P. Shu, et al., Chem. Eng. J. 454 (2023) 140053. doi: 10.1016/j.cej.2022.140053

    88. [88]

      Y.X. Shi, L.L. Li, Z. Xu, et al., Chem. Eng. J. 459 (2023) 141549. doi: 10.1016/j.cej.2023.141549

    89. [89]

      R.Z. Xiong, X.X. Ke, W.F. Jia, et al., J. Mater. Chem. A 11 (2023) 2178–2190. doi: 10.1039/d2ta09255g

    90. [90]

      Y.C. Wang, M.J. Liu, C.X. Wu, et al., Small 18 (2022) 2202544. doi: 10.1002/smll.202202544

    91. [91]

      Y.W. Xiao, B. Yao, M.H. Cao, et al., Small 19 (2023) 2207499. doi: 10.1002/smll.202207499

    92. [92]

      L.H. Xia, K.S. Zhang, X.D. Wang, et al., Appl. Catal. B: Environ. 325 (2023) 122387. doi: 10.1016/j.apcatb.2023.122387

    93. [93]

      H.Y. Liu, F. Sun, X.Y. Li, et al., Composites Part B: Eng. 259 (2023) 110746. doi: 10.1016/j.compositesb.2023.110746

    94. [94]

      Z.W. Zhao, Z.L. Wang, J.F. Zhang, et al., Adv. Funct. Mater. 33 (2023) 2214470. doi: 10.1002/adfm.202214470

    95. [95]

      H.B. Yu, J.H. Huang, L.B. Jiang, et al., Appl. Catal. B: Environ. 298 (2021) 120618. doi: 10.1016/j.apcatb.2021.120618

    96. [96]

      Y.H. Peng, X.L. Guo, S.F. Xu, et al., J. Energy Chem. 75 (2022) 276–284. doi: 10.1016/j.jechem.2022.06.027

    97. [97]

      H.F. Li, H.T. Yu, X. Quan, et al., Adv. Funct. Mater. 25 (2015) 3074–3080. doi: 10.1002/adfm.201500521

    98. [98]

      J.J. Lu, Y.H. Chen, L.N. Li, et al., Chem. Eng. J. 362 (2019) 1–11. doi: 10.1016/j.cej.2018.12.130

    99. [99]

      L.L. Wang, T. Yang, L.J. Peng, et al., Chin. J. Catal. 43 (2022) 2720–2731. doi: 10.1016/S1872-2067(22)64133-0

    100. [100]

      J.X. Bai, W.L. Chen, L. Hao, et al., Chem. Eng. J. 447 (2022) 137488. doi: 10.1016/j.cej.2022.137488

    101. [101]

      Y. Yuan, R.T. Guo, L.F. Hong, et al., Chemosphere 287 (2022) 132241. doi: 10.1016/j.chemosphere.2021.132241

    102. [102]

      Z.J. Chen, H. Guo, H.Y. Liu, et al., Chem. Eng. J. 438 (2022) 135471. doi: 10.1016/j.cej.2022.135471

  • Figure 1  Illustration of (A) Schottky junction, (B) ZnIn2S4-graphene junction, (C) type-Ⅰ junction, (D) type-Ⅱ junction, (E) p-n junction, (F) indirect Z-scheme junction, (G) direct Z-scheme junction, (H) S-scheme junction, (I) ternary junction, and (J) quaternary junction. RC and OC refer to reduction cocatalyst and oxidation cocatalyst, respectively. Copied with permission [48]. Copyright 2022, EcoMat.

    Figure 2  The improvement methods of ZnIn2S4-based S-scheme heterojunction.

    Figure 3  Charge-transfer processes in an S-scheme heterojunction. (A) before contact, (B) after contact, (C) photogenerated charge carrier transfer process in S-scheme mode. Copied with permission [51]. Copyright 2020, Elsevier lnc.

    Figure 4  (A) Atomic force microscopy image of CP composite, (B, C) corresponding surface potential distribution of CP in darkness (B) and under light irradiation (C). (D) The line-scanning surface potential from point A to B. (E) The schematic illustration of photoirradiation KPFM. Copied with permission [64]. Copyright 2021, Wiley-VCH GmbH.

    Figure 5  (A) Band structures of samples. Electron paramagnetic resonance (EPR) spectra of (B) DMPO-OH adducts (in aqueous solution) and (C) DMPO-O2 adducts (in methanol solution) over TiO2, ZnIn2S4, and TiO2@ZnIn2S4 under illumination for 2 min. Schematic diagrams of (D) conventional type-Ⅱ and (E) S-scheme heterojunction. Reproduced with permission [66]. Copyright 2021, Wiley-VCH GmbH.

    Table 1.  Summary table of the improvement methods of ZnIn2S4-based S-scheme heterojunction.

    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  1
  • 文章访问数:  66
  • HTML全文浏览量:  4
文章相关
  • 发布日期:  2025-04-15
  • 收稿日期:  2024-03-17
  • 接受日期:  2024-06-11
  • 修回日期:  2024-04-25
  • 网络出版日期:  2024-06-12
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章