Advances in Z-scheme and S-scheme heterojunctions for photocatalytic and photoelectrocatalytic H2O2 production

Xibao Li Yiyang Wan Fang Deng Yingtang Zhou Pinghua Chen Fan Dong Jizhou Jiang

Citation:  Xibao Li, Yiyang Wan, Fang Deng, Yingtang Zhou, Pinghua Chen, Fan Dong, Jizhou Jiang. Advances in Z-scheme and S-scheme heterojunctions for photocatalytic and photoelectrocatalytic H2O2 production[J]. Chinese Chemical Letters, 2025, 36(10): 111418. doi: 10.1016/j.cclet.2025.111418 shu

Advances in Z-scheme and S-scheme heterojunctions for photocatalytic and photoelectrocatalytic H2O2 production

English

  • H2O2 is an important efficient and versatile oxidizer and disinfectant, which has been widely used in chemical [1-5], medical [6-10], energy [11-15], and environmental fields [16-20] since its first synthesis by in 1818. Currently, commercially available H2O2 is mainly produced by electrolysis [21-25], conventional anthraquinone method [26-30], alcohol oxidation [31-35], cathodic reduction [36-40], and direct hydroxide oxidation [41-45]. Among them, anthraquinone method is the main method for industrial production of H2O2, which can produce H2O2 in large quantities, and its process is relatively stable and not easy to produce side reactions, but the process method is energy-consuming and costly, and there are also certain safety risks. Hence, there exists an imperative and pressing need to develop technologies for the safe, efficient, and environmentally benign production of H2O2. In this context, the deployment of photocatalytic and photoelectrocatalytic technology for the direct production of H2O2 from renewable energy sources has emerged as a highly significant research avenue [46-50]. Over the preceding decade, a significant annual escalation in both the quantity of scholarly publications and the frequency of citations pertaining to the photocatalytic and photoelectrocatalytic H2O2 production has been observed (Fig. S1 in Supporting information).

    The key to photocatalytic and photoelectrocatalytic technology is to develop high-efficiency and stable photocatalysts. The photocatalyst and photoelectrocatalyst can capture photons from the sun and generate electron-hole pairs, which causes oxidation and reduction reactions [51-55]. H2 generation [56-60], N2 fixation [61-65], antibiotic degradation [66-70] are the key research content of today's society. The photocatalytic H2O2 production necessitates the utilization of solar energy as a direct energy source, precluding the need for any intermediate energy conversion processes [71-75]. The photocatalytic process primarily entails the absorption of photons, the subsequent separation of photogenerated charge carriers, and the interfacial transfer of these carriers to facilitate the requisite reactions on the semiconductor photocatalyst surface [76-79]. Since the inaugural instance of employing ZnO for the H2O2 production [80], there has been a surge of research interest over the past few decades, focusing on the development of high-performance catalysts for H2O2 production under illumination conditions. The development and engineering of highly active and selective photocatalysts, achieved through the modulation of their intrinsic properties, have yielded a plethora of remarkable outcomes. Despite significant advancements, the solar-to-chemical conversion (SCC) efficiency of the photocatalysts remains unsatisfactory, with a maximum efficiency of only 10.1% [81]. Furthermore, the pinnacle of the apparent quantum efficiency (AQE) for non-sacrificial H2O2 production has reached 1.19% [82]. The principal challenges that constrain the efficiency of H2O2 production pertain to the limited light absorption range, swift charge recombination, sluggish surface kinetics for water oxidation, and diminished selectivity for the reduction of O2 [83-85].

    Photocatalysts and photoelectrocatalysts are generally in powder form, belonging to the nano-scale, dispersed in water, which is difficult to be recycled and reutilized, creating unnecessary waste in the experimental process. This realization has spurred researchers to explore more efficacious strategies aimed at enhancing the catalytic efficiency and optimizing the utilization of photocatalysts [86-88]. After continuous exploration, research workers found that the catalyst can be fixed on the substrate, and by applying an electric field, the photogenerated charges can be separated with greater efficiency under the synergistic effect of the internal and external electric fields, and the photogenerated electrons can be introduced into the counter electrode through the external circuit, weakening the photocatalytic catalyst charge compounding situation, and then an electrochemistry-assisted photocatalytic technology was developed [89-91], photoelectrocatalytic technology [92-95]. When light fields are introduced into conventional electrocatalytic techniques, the electronic properties of specific electrocatalysts are susceptible to disturbances, such as electron transfer, energy band bending, Fermi energy levels, and intermediate desorption energies, which can significantly alter the intrinsic pathways and performance of catalysis. Compared with monoelectrocatalysis, photoelectrocatalysis has higher energy efficiency and potential to activate small molecules at low overpotentials. Photoelectrochemical reactions are photomagnetization processes that generate photogenerated electrons-holes on semiconductor surfaces in contact with an electrolyte [96-100]. The electrons and holes are segregated on the catalyst surface, facilitating the reaction of electrons with oxygen to form H2O2, while holes react with water in a concurrent process. Photoelectrochemical (PEC) cells are capable of spatially segregating the electrolyte via two distinct half-channels, thereby mitigating the decomposition of H2O2. In addition, photo/electrocatalytic can also be used to degrade HCHO in indoor air [101].

    Among the various modification methods, the creation of a suitable heterojunction structure was considered an effective strategy. Shen et al. [102] successfully engineered a high-low junction photocatalyst by integrating BiOBr with Bi2S3, characterized by an extensive interfacial contact area. This integration was achieved through a synergistic approach combining the molten salt method with an ion exchange strategy. Specifically, Bi2S3 was epitaxially grown on the surface of BiOBr via a high-temperature molten salt reaction, resulting in an intimate interface between the two materials. The deep VB position of BiOBr, coupled with the narrow band gap of Bi2S3, induces an intrinsic internal electric field (IEF) and band bending at the heterojunction. This configuration effectively facilitates the separation and transfer of photogenerated charge carriers, thereby enhancing the photocatalytic performance. Moreover, BiOBr/Bi2S3 retains a high oxidation potential, which significantly boosts its photocatalytic oxidation activity. Interestingly, in the molten state, the close binding between BiOBr and Bi2S3, facilitated by the ion-exchange strategy, promotes the in-situ H2O2 production, further augmenting the photocatalytic efficiency. Mechanisms such as Z-scheme and S-scheme are the most common ones used in the literature to characterize charge transfer in heterojunction structures. The concept of Z-scheme heterojunction was initially proposed by Bard in 1979 and has since been widely adopted in the field of photocatalysis. Cheng et al. [103] synthesized Z-scheme Ag/ZnFe2O4-Ag-Ag3PO4 composites to produce H2O2 by two consecutive steps of single-electron oxygen reduction. In recent years, Z-scheme heterojunction was widely used for photocatalytic CO2 reduction [104-108], H2 generation [109-113], N2 fixation [114-118] and H2O2 production [119-123].

    However, there is still some confusion about the mechanism of Z-scheme heterojunction. S-scheme heterojunction is composed of a reduced semiconductor and an oxidized semiconductor, with the oxidized component being capable of existing as either a p-type or an n-type semiconductor. The migration of photogenerated charge carriers is effectively facilitated by the IEF present at the interfaces of the disparate semiconductors, thereby sustaining a high redox potential [124-127]. Li et al. [128] employed a special self-assembly strategy to fabricate an innovative S-scheme composite by integrating sulfur-doping porous graphite carbon nitride (S-pCN) with WO2.72. During the formation of the heterojunction, the electron-deficient state associated with surface oxygen vacancies on WO2.72 was alleviated by the lone pair electrons of S and N elements present in S-pCN. By meticulously adjusting the composite ratios, an optimal number of surface oxygen vacancies were retained on WO2.72, which in turn facilitates electron migration and the generation of free radicals. The IEF and band bending effects at the interface accelerate the transfer of photogenerated charges, thereby promoting the separation of less reactive photogenerated carriers and preserving those with higher redox potentials, specifically electrons (e) and holes (h+). In recent years, S-scheme heterojunction has received unprecedented attention for their excellent photocatalytic activity. They are widely used for photocatalytic CO2 reduction [129-133], H2 generation [134-138], N2 fixation [139-143], antibiotic degradation [144-148] and H2O2 production [149-153].

    As delineated in the abstract of the article depicted in Fig. 1, in this review, we commenced with a discussion of the mechanisms associated with Z-scheme and S-scheme heterojunctions within the realm of photocatalytic and photoelectrocatalytic H2O2 production. Subsequently, we introduced in detail the preparation strategies and characterization techniques employed for the development of Z-scheme or S-scheme photocatalysts and photoelectrocatalysts for the H2O2 production, and briefly described the multifunctional applications of such catalysts in the H2O2 production. Ultimately, the article concluded with a prospective outlook on future research directions and the unresolved issues that warrant attention.

    Figure 1

    Figure 1.  Summary of this review content.

    The H2O2 production from semiconductor photocatalysts driven by light irradiation encompasses a series of intricate and sequential reaction processes [154-157], as shown in Fig. 2. Initially, upon irradiation of the photocatalyst with light of sufficient energy, electrons within the valence band (VB) are excited and transition to the conduction band (CB), thereby generating holes (h+) in the VB. The photogenerated electrons, along with hydrogen ions, migrate toward the surface of the photocatalyst. During this migration, charge carriers are prone to recombine, resulting in the dissipation of energy as light or heat. The surviving carriers that reach the catalyst surface participate in reduction or oxidation reactions with molecular O2 and aqueous H2O, respectively, while a portion of them inevitably undergo recombination [158-160]. Ultimately, the spatially separated electrons and holes participate in surface redox reactions, primarily the oxygen reduction reaction (ORR) and the water oxidation reaction (WOR), thereby culminating in the H2O2 production [161-163]. Hence, to enhance the H2O2 production, it is imperative to retard the recombination of photogenerated e and h+, while concurrently ensuring that the photocatalytic system possesses a robust redox potential capable of facilitating H2O2 production. Typically, the photocatalytic H2O2 production is predominantly achieved through the proton-coupled two-electron ORR, which can also proceed either through a sequential one-electron pathway (Fig. 2, Eqs. 1 and 2) or via a coordinated two-electron pathway (Fig. 2, Eq. 3) [164-166].

    Figure 2

    Figure 2.  Schematic diagram of relevant mechanisms of the photocatalytic H2O2 production.

    In the case of the former pathway, where H2O2 is synthesized via a two-step electron reduction coupled with a two-step hole oxidation process, the rate-determining step is the formation of the superoxide radical (O2) intermediate. This step necessitates a potential that is more negative than −0.33 V relative to the normal hydrogen electrode (NHE), as indicated in Eq. 1. Conversely, in the latter pathway, this intermediate formation is bypassed, as O2 is directly reduced to H2O2 in a single, concerted step. Consequently, a synergistic two-electron pathway for ORR is chemically more favorable. Additionally, the oxidation of H2O2 via a two-electron WOR has been documented. Nevertheless, this process is fraught with challenges, primarily due to the high standard potential of +2.73 V vs. NHE, the sluggish kinetics involved, and the relatively low selectivity when compared to the O2 evolution process from H2O. Despite the significant potential of photocatalytic H2O2 production, the yield of H2O2 remains unsatisfactory, attributed to rapid recombination of photogenerated e and h+, as well as the relatively slow kinetics of the redox reactions involved. Furthermore, in the context of single-component photocatalysts, an inherent trade-off exists between redox capacity and the spectral range of light absorption. The construction of step-structured (S-scheme) heterojunction has emerged as an innovative strategy to mitigate these challenges [167-170].

    Two-step electron reduction channel (Eqs. 1–3):

    O2+eO2(0.33V)

    (1)

    O2+H+OOH(0.13V)

    (2)

    OOH+H++eH2O2(0.046V)

    (3)

    Two-step hole oxidation channel (Eq. 4):

    h++OHOH+OHH2O2(2.29V)

    (4)

    Z-scheme photocatalysts have garnered significant attention within the realm of photocatalysis due to their unique electronic configurations and enhanced performance in various energy conversion and environmental remediation applications. These photocatalysts can be broadly categorized into two main types: direct and indirect Z-scheme photocatalysts. Direct Z-scheme photocatalysts consist of two semiconductor materials with overlapping bandgaps, which facilitate the transfer of photogenerated e and h+. In contrast, indirect Z-scheme photocatalysts utilize a redox mediator to bridge the electron transfer gap between the two semiconductors. The redox mediator serves as an electron acceptor for one semiconductor and an electron donor for the other, thereby establishing a cyclic electron transfer mechanism. This cyclic pathway enables sustained photocatalytic activity and enhances the rate of H2O2 production [171-175].

    The S-scheme was initially conceived as a remedial approach to address the inherent challenges associated with Type Ⅱ or Z-scheme photocatalysts [176-180], and has recently been successfully applied to a wide range of critical photocatalytic reactions, including H2 release [181-185], CO2 reduction [186-190], pollutant degradation [191-195], N2 immobilization [196-200], and H2O2 production [201-205]. At the core, S-scheme systems are constituted by two intimately associated semiconductor photocatalysts with interdigitated energy band diagrams. The photocatalyst characterized by a higher conduction band minimum (CBM), valence band maximum (VBM), and Fermi energy level is designated as the reduced photocatalyst (RP), whereas its counterpart is referred to as the oxidized photocatalyst (OP). Owing to the higher Fermi energy level of the RP compared to that of the OP, free electrons spontaneously migrate from the RP to the OP upon intimate contact, thereby equilibrating the Fermi energy levels at the interface. This electron transfer results in a net positive charge on the RP side and a negative charge on the OP side at the interface, thereby establishing an IEF oriented from the RP to the OP. Upon light irradiation, when electron-hole pairs are generated within both photocatalysts, the IEF facilitates the migration of photogenerated electrons from the OP's CB to the RP's VB, where they recombine with the photogenerated holes. Concurrently, the CB electrons in the RP and VB holes in the OP migrate towards the surface of the photocatalysts to participate in catalytic reactions. These charge carriers, endowed with maximal redox capacity, provide a more robust thermodynamic driving force for photocatalytic H2O2 production. Consequently, the S-scheme heterojunction not only enhances the transfer and separation of charge carriers but also accelerates the redox reactions, culminating in the efficient photocatalytic H2O2 production [206-210].

    Photoelectrocatalytic H2O2 production is usually carried out by ORR or WOR, as shown in Fig. S2 (Supporting information) and the production of H2O2 by semiconductor photoelectrocatalytic materials, for example, involves the following processes. Initially, the photocathode within the photocatalytic system absorbs photons with energy exceeding its bandgap (Eg), thereby generating electron-hole pairs. Subsequently, the applied potential promotes the transfer of electrons to activated oxygen species, thereby inducing the spatial separation of photogenerated e and h+ pairs [211-213]. These charge carriers migrate to the CB and VB of the semiconductor, respectively, where they participate in a variety of redox reactions. It is worth noting that the light absorption capacity of a semiconductor is directly proportional to its band gap width; a larger band gap requires higher-energy photons to induce electronic excitation. Moreover, the energy band structure of a semiconductor is intricately connected to the redox reactions occurring on its surface. Therefore, in order to achieve efficient photocatalytic H2O2 production, the energy band positions of the applied semiconductors need to be coordinated to achieve high yields.

    There are multiple electron pathway reactions in the ORR process, such as one-electron reduction to O2, two-electron reduction to H2O2, and four-electron reduction to H2O. Kinetically, O2 would be more inclined to undergo a four-electron reaction to produce H2O. Currently, it is a widely accepted notion that H2O2 can be produced via either a two-step one-electron indirect reduction pathway or a one-step two-electron direct reduction pathway. Han et al. synthesized CoSe2 with varying Se stoichiometry using a straightforward hydrothermal approach [214]. They induced a phase transition of CoSe2 from the orthorhombic to the cubic phase, selectively for Se-rich and Se-deficient compositions, by incorporating NaBH4. The resulting catalyst demonstrated exceptional activity, high selectivity for H2O2, and enduring stability during a two-electron acidic ORR in 0.1 mol/L HClO4. The yield of H2O2, ascertained through a flow cell configuration, amounted to 115.92 mmol gcat-1 h-1, thereby underscoring its prospective viability for deployment in wastewater treatment endeavors. This investigation markedly augmented the comprehension of the crucial function that catalyst crystal phase and defect engineering played in bolstering the catalytic efficacy of the two-electron ORR towards H2O2 production. Alternatively, solar synthesis of H2O2 can be via the 2e route from ORR, which is achieved by applying a photocatalytic system or a photocathode in a PEC system. In general, the efficacy of a photoelectrocatalytic system for H2O2 production is contingent upon several pivotal factors, including the morphology and electronic energy band structure of the catalyst utilized, in addition to the composition of the reaction solution (Eqs. 5–7) [215-217].

    O2+eO2E0=0.33V

    (5)

    O2+2H+H2O2E0=1.44V

    (6)

    O2+2H++2eH2O2E0=0.68V

    (7)

    The production of H2O2 by the photoelectrocatalytic WOR pathway is a two-electron process. The semiconductor photoanode is fixed in the middle of the electrolyte through an electrode fixture and is in full contact with the electrolyte. Upon the incidence of photons whose energy equals or surpasses the band gap of the photoanode, electrons within the VB of the semiconductor are promoted to the conduction band, thereby generating photogenerated holes within the VB. These photogenerated holes migrate to the surface of the photoanode, which possesses potent oxidizing capabilities, facilitating the oxidation of H2O to H2O2 at the electrolyte-photoanode interface. Concurrently, the photogenerated electrons are conveyed to the counter electrode via the external circuit under the influence of an applied bias voltage, where they reduce H+ to H2. H2O2 is produced at the photoanode through H2O oxidation, however, the production and accumulation of H2O2 by H2O oxidation needs a more positive than the production of O2 from H2O potential. Therefore, it does not occur readily (Eqs. 8–10) [218-220].

    2H2O+2h+H2O2+2H+E0=1.78V

    (8)

    H2O+h+OH+H+E0=2.73V

    (9)

    OH+OHH2O2E0=1.14V

    (10)

    The fabrication of Z-scheme and S-scheme heterojunction catalysts represents a crucial approach for the efficient photocatalytic and photoelectrocatalytic H2O2 production. A scientific and stable interfacial structure can effectively facilitate separation and migration of photogenerated charge carriers, thereby significantly improving the efficiency of H2O2 production via photocatalytic pathways. Consequently, reliable preparation strategies for heterojunctions are of particular importance. The successful preparation of Z-scheme and S-scheme heterojunction catalysts with high-quality interfacial morphology commonly employs several strategies, including in-situ growth, self-assembly, deposition, one-step calcination, ion exchange, Interlayer coordination, template method (Fig. S3 in Supporting information). The following discussion will delve into these preparation strategies in detail.

    The in-situ growth method refers to a strategy whereby a second semiconductor nucleates and grows on the surface or interface of a pre-synthesized semiconductor under specific conditions, thereby achieving a close integration of the two semiconductors to form a heterojunction. The advantage of the in-situ growth strategy lies in its simple preparation process and the tight interfacial bonding of the formed heterojunction. However, its disadvantage is that the in-situ synthesis of multi-component heterojunctions may lead to uneven component distribution due to differences in reaction kinetics. Various successfully synthesized Z-scheme or S-scheme heterojunction catalysts for the photocatalytic H2O2 production, such as CdS/RF [46], TiO2@BTTA [201], iCOF/BO [203], ZnO/PBD [221], WS2/g-C3N4 [119], ZIS-Z/OCN [122], r-CNW [121], CdS/Bi2WO6 [153], BiOBr/WO3 [120], have been fabricated using in-situ growth strategies, including methods like oil bath, photodepositing, hydrothermal and solvothermal.

    Specifically, Zhu et al. employed the oil bath method to fabricate a CdS/RF S-scheme heterojunction composite by in-situ growth of CdS nanoparticles on the surface of resorcinol-formaldehyde (RF) resin spheres (Fig. 3a) [46]. The Field-emission scanning electron microscopy (FESEM) images clearly demonstrated the uniform growth of CdS nanoparticles on the surface of RF spheres. It is noteworthy that during the synthesis of CdS/RF, sodium citrate was found to be an essential chelating agent for the formation of the core-shell structure. In the absence of sodium citrate, CdS nanoparticles were randomly distributed within the composite rather than forming a uniform shell on the surface of the RF spheres. The successfully synthesized CdS/RF S-scheme heterojunction composite achieved a H2O2 release rate of up to 801 µmol L-1 h-1 under simulated solar irradiation. Moreover, Xia et al. utilized the photodeposition method combined with the oil bath method to in-situ grow Bi2O3(BO) nanoparticles on the ionic covalent organic framework (iCOF) nanofibers, successfully constructing the iCOF/BO S-scheme heterojunction composite for photocatalytic H2O2 production via a dual-channel pathway (Fig. 3b) [203]. The FESEM, TEM, HRTEM, high-angle annular dark-field (HAADF) images, and the corresponding elemental maps substantiated the uniform distribution of BO nanoparticles on the surface of iCOF nanofibers, providing further evidence for the successful preparation of the iCOF/BO composite. Under oxygen-saturated and visible light irradiation conditions, the synthesized iCOF/BO S-scheme heterojunction composite exhibits a H2O2 production efficiency of 9.76 mmol g-1 h-1 in pure water. Recently, Yu et al. placed a mixture containing a certain amount of TiO2 nanofibers, acetic acid, acetonitrile, 4,4′,4′′-(1,3,5-triazine-2,4,6-triyl)triphenylamine (TA), and 1,3,5-benzenetricarbaldehyde (BT) after ultrasonication into a Pyrex tube, filled it with N2, and maintained it at room temperature for 72 h [201]. They successfully grew a covalent organic framework (labeled as BTTA) synthesized through a Schiff base condensation reaction in situ on the surface of TiO2 nanofibers (TiO2 NFs), resulting in the preparation of a TiO2 @BTTA S-scheme heterojunction composite material (Fig. 3c). The FESEM and high-resolution transmission electron microscopy (HRTEM) images, along with energy-dispersive X-ray (EDX) spectra, substantiated the uniform growth of BTTA on the surface of TiO2 NFs. The porosity and encapsulation capability of BTTA endowed the composite material with a wealth of active sites and superior light absorption performance. The resulting TiO2 @BTTA composite material achieved a H2O2 release rate of 740 µmol L-1 h-1 and a 96% conversion rate of furfural alcohol under light irradiation.

    Figure 3

    Figure 3.  (a) Preparation route of CdS/RF composite. Copied with permission [46]. Copyright 2023, Elsevier. (b) The synthesis process for iCOF/BO composite. Copied with permission [203]. Copyright 2024, Editorial Office of Acta Physico-Chimica Sinica. (c) Synthetic procedure of TiO2@BTTA composites. Copied with permission [201]. Copyright 2023, Elsevier.

    The solvothermal or hydrothermal method is one of the commonly used approaches to implement the in-situ growth strategy for constructing S-scheme or Z-scheme heterojunction catalysts for photocatalytic H2O2 production. The hydrothermal method is a process in which, within a sealed pressure vessel, using water as the solvent, a high-temperature and high-pressure reaction environment is created by heating and pressurizing, enabling substances that are usually insoluble or sparingly soluble to dissolve and recrystallize. For instance, Wu et al. employed the Suzuki-Miyaura reaction and the hydrothermal method [221]. With 1,6-bis(pyrene)-benzodithiophene serving as the electron donor and 2,6-dibromobenzodithiophene-4,8-dione as the electron acceptor, 1,6-bis(pyrene)-benzodithiophene dione (PBD) was synthesized, and ZnO nanoparticles were in-situ grown on the PBD substrate, successfully fabricating the ZnO/PBD S-scheme heterojunction composite (Fig. 4a). Through transmission electron microscopy (TEM) and HRTEM images, it can be clearly observed that the ZnO nanoparticles, as highly crystalline single crystals, were uniformly distributed on the amorphous PBD nanosheets. Under oxygen-rich and illumination conditions, ZnO/PBD has a H2O2 production efficiency of 4.07 mmol g-1 h-1 in 10 vol% methanol solution. The solvothermal method is a material preparation method developed on the basis of the hydrothermal method. It mainly uses organic solvents or non-aqueous media as the reaction medium and promotes chemical reactions by controlling the temperature and pressure in a closed system. For instance, Li et al. utilized the solvothermal method to control the growth of WS2 on S-doped graphitic carbon nitride (SCN) and obtained excellent interfacial characteristics by means of horizontal implantation, thus preparing the WS2/SCN Z-scheme heterojunction composite (Fig. 4b) [119]. TEM and HRTEM images clearly showed that WS2 nanosheets grew horizontally and uniformly on the SCN nanosheets, demonstrating the successful construction of the WS2/SCN heterojunction. Under light illumination, the successfully prepared WS2/SCN Z-scheme heterojunction composite exhibited a H2O2 production rate of up to 5216 µmol L-1 h-1 in 10 vol% isopropanol solution. Moreover, Liu et al. utilized the solvothermal method to in-situ grow oxygen-doped graphitic carbon nitride (OCN) on Zn vacancy-containing ZnIn2S4 (ZIS-Z) flower-like microspheres, successfully fabricating a 2D/2D ZIS-Z/OCN Z-scheme composite (Fig. 4c) [122]. The SEM, atomic force microscopy (AFM), TEM, and HRTEM images clearly displayed that the ultrathin nanosheets OCN with a porous structure were highly dispersed on ZIS-Z with a flower-like microsphere structure composed of numerous ultrathin, smooth, and intertwined nanosheets. These observations further substantiated the intimate contact between ZIS-Z and OCN and provided comprehensive validation for the epitaxial formation of the ultrathin 2D/2D heterojunction in the ZIS-Z/OCN composite. Under visible light irradiation for 80 min, the removal rate of chlortetracycline hydrochloride (CTC) and the production rate of H2O2 by ZIS-Z/OCN were 88.43% and 168.67 µmol/L, respectively.

    Figure 4

    Figure 4.  (a) Synthesis route of ZnO/PBD. Copied with permission [221]. Copyright 2024, Editorial Office of Acta Physico-Chimica Sinica. (b) Formation illustration of WS2/S-g-C3N4 Z-scheme heterostructures. Copied with permission [119]. Copyright 2024, Elsevier. (c) Schematic representation of preparing ZIS-Z/OCN. Copied with permission [122]. Copyright 2023, Elsevier.

    The self-assembly method is an approach which enables two or more semiconductors to spontaneously arrange and assemble into a stable structure based on the interactions of non-covalent bonds. These non-covalent bond interactions in the preparation of Z-scheme or S-scheme heterojunction catalysts for H2O2 production can be mainly classified into electrostatic, π-π stacking, and hydrogen bonding interaction. Among them, the latter two interactions generally occur simultaneously to assist two semiconductors in self-assembling and thus closely combining to form a heterojunction. The advantage of the self-assembly strategy lies in its ability to achieve precise design of the interface, and the preparation conditions conform to the principle of environmental friendliness. However, it is also prone to the drawback of insufficient structural stability.

    Self-assembly guided by electrostatic interactions can be mainly divided into two categories. One is dominated by ionic bonding. For example, in Z-scheme or S-scheme heterojunction catalysts such as ZnO/COF(TpPa-Cl) [222], ZnSe QD/COF [151], OCN@CNT-x [223], the semiconductors with opposite charges are attracted to each other by electrostatic attraction, thus forming stable heterojunctions. The other is dominated by charge transfer. For instance, in Z-scheme or S-scheme heterojunction catalysts like 3DOM SCN/T [224], BP/BiOBr [225], Bi4Ti3O12/g-C3N4 [226], charge transfer occurs between distinct semiconductors. This process not only facilitates the self-assembly of the composite but also results in the formation of stable heterojunctions.

    Electrostatic self-assembly dominated by ionic bonding usually requires adjusting the pH values of semiconductor solutions to achieve opposite charge properties between the two semiconductors. Specifically, Zhang et al. combined positively charged ZnO with negatively charged TpPa-Cl to prepare the ZnO/TpPa-Cl (ZT) S-scheme heterojunction catalyst (Fig. 5a) [222]. By measuring the zeta potentials of ZnO and TpPa-Cl respectively, the realization of the electrostatic self-assembly process dominated by ionic bonding was theoretically supported. At a pH value of 7, ZnO nanoparticles exhibited a positive zeta potential, whereas TpPa-Cl showed the opposite. Consequently, ZnO and TpPa-Cl could be combined together through electrostatic attraction, which was conducive to the formation of a stable heterojunction. Through TEM and HRTEM images, the close combination of ZnO and TpPa-Cl could be clearly observed, further demonstrating the successful preparation of the ZT heterojunction composite. Under oxygen-rich and simulated solar light irradiation conditions, ZT has the maximum H2O2 production efficiency of 2443 µmol g-1 h-1 in 10 vol% ethanol solution.

    Figure 5

    Figure 5.  (a) Schematic illustration of the formation of the ZT heterojunction. Copied with permission [222]. Copyright 2022, Elsevier. (b) Schematic representation of the synthesis process for 3DOM SCN/T. Copied with permission [224]. Copyright 2023, Elsevier. (c) Schematic depiction of the preparation process for Sol-BP/BiOBr composites. Copied with permission [225]. Copyright 2020, Elsevier.

    The electrostatic self-assembly process dominated by charge transfer usually occurs under certain conditions. After two semiconductors are fully contacted through continuous stirring or ultrasonic treatment and other means, one semiconductor transfers electrons to the interface where it contacts the other semiconductor, thus driving the self-assembly process and forming a stable heterojunction. For example, Jiang et al. employed the electrostatic self-assembly strategy, which is primarily driven by charge transfer, to combine TiO2 with three-dimensionally ordered macroporous sulfur-doped graphitic carbon nitride (3DOM SCN), thereby fabricating the 3DOM SCN/TiO2 (SCN/T) S-scheme composite (Fig. 5b) [224]. Through comparing the differences in the binding energies of various elements in the XPS spectra between the single SCN substance and the SCN/T composite, it could be seen that when SCN and TiO2 were fully contacted, electrons migrate from the interface of SCN to TiO₂, thus driving the self-assembly process and forming a stable heterojunction. Through SEM and TEM images, it could be clearly observed that TiO2 nanoparticles were uniformly anchored on the SCN framework, further demonstrating the successful preparation of the SCN/T heterojunction composite. Under oxygen-rich and 300 W Xe lamp irradiation conditions, SCN/T has the maximum photocatalytic H2O2 production efficiency of 2128 µmol g-1 h-1 in deionized water. Furthermore, Li et al. also adopted the analogous strategy dominated by charge transfer to combine exfoliated black phosphorus (BP) nanosheets with BiOBr nanosheets possessing a relatively high Fermi level [225]. In particular, they proposed to use the ultrasonic-assisted solvothermal method instead of the conventional simple ultrasonic combination method to construct a novel layered BP/BiOBr S-scheme heterojunction with a more uniform and stable structure (Fig. 5c). By comparing the binding energies of various elements in the XPS spectra of pure BP, pure BiOBr and the BP/BiOBr composite, it could be found that when BP and BiOBr were in contact with each other, electrons migrate from the interface of BiOBr to BP, thus driving the self-assembly process and forming a stable heterojunction. Through the SEM and HRTEM images, the fine structures of the crystal planes and phase interfaces of the BP/BiOBr distinctive heterojunction could be clearly witnessed, further demonstrating the effective synthesis of the BP/BiOBr heterojunction. Under the conditions of oxygen-rich and irradiation with a 300 W Xe lamp (420 nm < λ < 780 nm), the efficiency of BP/BiOBr for producing H₂O₂ in pure water is 1.62 µmol L-1 min-1.

    π-π Stacking represents a unique spatial arrangement of aromatic compounds, denoting a weak interaction that frequently occurs between aromatic rings. This non-covalent interaction is as significant as hydrogen bonding. Besides the self-assembly driven by electrostatic interactions, π-π stacking or the synergistic effect of π-π stacking and hydrogen bonding usually also serves as a driving force to guide the self-assembly process of two semiconductors, facilitating the formation of stable heterojunctions at the semiconductor interfaces. For example, heterojunction catalysts such as PDI-Ala/S-C3N4 [227], CoPc/K/Na/PCN [228], Co-N-C/SA-PDI [229], PDI/g-C3N4 [230], were all driven by π-π stacking or the synergistic effect of π-π stacking and hydrogen bonding in the self-assembly process of two semiconductors, thus forming stable heterojunctions. In particular, Li et al. based on the fact that g-C3N4 has an aromatic heterocyclic structure and that N,N′-bis(propionic acid)-perylene-3,4,9,10-tetracarboxylic diimide (PDI-Ala) has a rigid planar structure with the characteristic of delocalized large π bonds, utilized π-π interactions to self-assemble PDI-Ala on the surface of sulfur-doped g-C3N4 (S-C3N4) nanosheets, thereby preparing PDI-Ala/S-C3N4 S-scheme composite (Fig. 6a) [227]. Through X-ray diffraction (XRD), infrared spectroscopy and DFT calculations, it was inferred that PDI-Ala self-assembles into supramolecules through lateral hydrogen bonds and longitudinal π-π stacking, while S-C3N4 and PDI-Ala were closely combined through π-π interactions and N—C bonds. Through TEM and elemental mapping, it was clearly observed that PDI-Ala is uniformly distributed on the porous two-dimensional S-C3N4 nanosheet structure, further demonstrating the successful preparation of the PDI-Ala/S-C3N4 heterojunction composite. Under the irradiation of a 300 W xenon lamp (420 nm < λ < 780 nm), the yield of H2O2 of the PDI-Ala/S-C3N4 in distilled water is 28.3 µmol g-1 h-1. Moreover, Liu et al. adopted a simple hydrothermal method. CoPc was self-assembled on the surface of alkali metal-doped g-C3N4 nanosheets (K/Na/PCN) through π-π stacking and hydrogen bonding interactions to prepare CoPc/K/Na/PCN Z-scheme composite (Fig. 6b) [228]. Based on the results of infrared (FTIR) spectroscopy, it could be analyzed and inferred that CoPc was combined with K/Na/PCN through non-covalent bonds, and the planar CoPc was introduced onto the surface of K/Na/PCN through π-π stacking and hydrogen bonding effected and covered the surface of K/Na/PCN nanosheets in the form of single or multiple molecular layers. As the amount of CoPc increased, the degree of self-aggregation of CoPc on the surface of K/Na/PCN also increased. Through the SEM and elemental mapping images, it can be clearly observed that CoPc uniformly and stably covers the surface of K/Na/PCN nanosheets, further demonstrating the successful preparation of CoPc/K/Na/PCN heterojunction composite. Under the conditions of oxygen-saturated adsorption and irradiation with a 60 W blue LED bulb, CoPc/K/Na/PCN can oxidize amines into imines in an ethanol solution with a certain amount of amines added, and simultaneously couple to produce H2O2 with a yield ranging from 3.27 mmol g-1 h-1 to 3.40 mmol g-1 h-1.

    Figure 6

    Figure 6.  (a) The fabrication process of PDI-Ala/S-C3N4. Copied with permission [227]. Copyright 2021, Editorial Office of Acta Physico-Chimica Sinica. (b) Synthetic pathways of CoPc/K/Na/PCN. Copied with permission [228]. Copyright 2024, Elsevier.

    The deposition method is one of the more common strategies for constructing Z-scheme or S-scheme heterojunction catalysts, apart from the in-situ growth and self-assembly methods. The deposition strategy generally refers to a method of preparing heterojunction composite materials by depositing one semiconductor directly onto the surface of another semiconductor in a certain way. This deposition process usually occurs in a solution. Depending on the differences in the environment during the solution stirring, we classify the deposition strategy into the in-situ deposition method and the chemical bath deposition method here. The in-situ deposition method usually involves the solution stirring process at room temperature, while the chemical bath deposition method usually conducts the solution stirring under the conditions of water bath or chemical bath heating. The advantage of the deposition strategy is its wide applicability and the potential for large-scale preparation of catalysts. However, it also has the disadvantage of a relatively long preparation process.

    Specifically, Meng et al. adopted the in-situ deposition strategy to deposit g-C3N4 nanosheets in situ onto the crystalline material (ZIF-L) formed by zinc atoms and the organic ligand 2-methylimidazole (2-MI) through coordination bonds, thus preparing ZIF-L/g-C3N4 Z-scheme heterojunction composite materials (ZCx) for the efficient production of H2O2 (Fig. 7a) [231]. Through the SEM and HRTEM images, it can be clearly observed that the fragmented and laminated g-C3N4 nanosheets were evenly distributed on the surface of ZIF-L with a smooth leaf-like morphology, and the two were closely combined, which demonstrates the successful preparation of the ZIF-L/g-C3N4 heterojunction composite. Under the irradiation condition with a 300 W xenon lamp, continuous ultrasonication of ZIF-L/g-C3N4 in ultrapure water can achieve a H2O2 yield of 1.45 mmol g-1 h-1.

    Figure 7

    Figure 7.  (a) Schematic representation of the fundamental synthetic protocol for ZCx. Copied with permission [231]. Copyright 2024, Elsevier. (b) Schematic illustration of the synthesis process for ZnO/ZnIn2S4. Copied with permission [232]. Copyright 2023, Elsevier. (c) Schematics for the synthesis of ZnO/CuInS2 photocatalyst. Copied with permission [50]. Copyright 2024, Wiley-VCH.

    Wu et al. adopted the electrospinning and chemical bath deposition methods to deposit ZnIn2S4 nanoparticles onto ZnO fibers, thus preparing a low-dimensional ZnO/ZnIn2S4 S-scheme photocatalyst for the efficient production of H2O2 (Fig. 7b) [232]. Through FESEM, TEM, HRTEM, HADDF images, as well as EDX elemental mapping results, it was clearly observed that several ZnIn2S4 nanoparticles were uniformly loaded on the surface of ZnO, and the composite exhibited a core-shell architecture, with ZnO constituting the core and ZnIn2S4 forming the shell. Concurrently, ZnIn2S4 nanoparticles were predominantly synthesized on the matrix of ZnO nanofibers, thereby evidencing the intimate contact between ZnO and ZnIn2S4 and further corroborating the successful fabrication of the ZnO/ZnIn2S4 heterojunction composite. Under the conditions of oxygen saturation and simulated sunlight irradiation, the yield of H2O2 of ZnO/ZnIn2S4 in 10 vol% methanol solution could reach 928 µmol g-1 h-1.

    Meng et al. prepared a plasmonic near-infrared-responsive ZnO/CuInS2 S-scheme heterojunction photocatalyst for the photocatalytic oxidation of glycerol to produce H2O2 through an in-situ deposition strategy [50]. Specifically, a certain amount of ZnO nanocages were uniformly dispersed in n-hexane. Subsequently, a certain amount of CuInS2 quantum dot (QD) solution was rapidly added under vigorous stirring. Finally, the ZnO/CuInS2 quantum dot heterostructure was obtained by vacuum-drying (Fig. 7c). The FESEM, STEM, HRTEM, HADDF images, and EDX elemental mappings clearly showed that ZnO/CuInS2 had a uniform particle size distribution, a distinct cubic morphology, and a hollow internal structure, further confirming the successful preparation of the composite. Upon the introduction of 10 mg of ZnO/CuInS2 into a mixed solution comprising 30 mL of deionized water and 4.5 µL of glycerol, following the attainment of adsorption equilibrium of oxygen within the solution, the mixture was subjected to irradiation from a 300 W Xe lamp (wavelength range: 350–1100 nm), after 4 h, the glycerol oxidation rate of the ZnO/CuInS2 composite reached a maximum of 0.27 mmol L-1 h-1, and correspondingly, the photocatalytic H2O2 production rate was 0.31 mmol L-1 h-1. In addition, Pradhan et al. [124] employed an in-situ deposition strategy and utilized a one-pot solvothermal method to in-situ deposit Bi2MoO6 nanosheets onto InVO4/CeVO4, thereby constructing a series of Bi2MoO6/InVO4/CeVO4 ternary S-scheme heterostructures for highly efficient photocatalytic production of H2O2 and H2. Among them, the yield of H2O2 could reach 1700 µmol L-1 g-1 h-1. The ternary heterostructure exhibited a unique morphology composed of Bi2MoO6 nanoplates, InVO4 nanorods, and CeVO4 nanosheets. Moreover, Xu et al. also prepared a graphitic carbon nitride/polyhedral oligomeric silsesquioxane PDI (p-CN/P-PDI) S-scheme composite through in-situ deposition strategy [55].

    Among the strategies for preparing heterojunction catalysts used in photocatalytic production of H2O2, in addition to the commonly used in-situ growth, self-assembly and deposition methods, several other preparation strategies have also been mentioned in some research papers, such as one-step calcination, ion exchange, interlayer coordination, template method and machine learning. These have provided novel and referable preparation ideas for researchers in the field of preparing Z-scheme or S-scheme heterojunction composites for photocatalytic production of H2O2. Herein, we will briefly introduce several other preparation strategies, aiming to provide references for researchers and help them explore better preparation strategies.

    Specifically, Zhou et al. [233] adopted a straightforward and expeditious one-step calcination approach. Initially, they homogeneously blended titanate powder with a urea solution of specified concentration and subjected the mixture to drying at a predetermined temperature. Subsequently, the dried material was directly subjected to high-temperature calcination, yielding mesoporous TiO2-x/g-C3N4 S-scheme composite that were abundant in oxygen vacancies (Fig. 8a). Chen et al. [234] adopted the ion exchange strategy and prepared the hollow tubular bifunctional In2S3/MnIn2S4 S-scheme heterojunction photoelectrocatalyst by subjecting In2S3 to an exchange treatment with Mn2+. Specifically, In2S3 was thoroughly dispersed in ethanol, followed by the rapid addition of a Mn2+ ethanol solution of a specific concentration. The resultant mixture was stirred at 60 ℃ for 2 h, then filtered and washed with ethanol and deionized water. Finally, the product was dried under vacuum to obtain the In2S3/MnIn2S4 composite (Fig. 8b). Liu et al. [127] adopted the strategies of interlayer coordination and interlayer separation to prepare a series of graphitic carbon nitride/polyoxometalate (CN-HMoP) S-scheme heterojunction photocatalysts. Specifically, a predetermined quantity of phosphomolybdic acid was dissolved in a ternary solvent mixture composed of H2O, ethanol, and DMF in a volumetric ratio of 3:2:1. Graphitic carbon nitride (CN) was added and then subjected to ultrasonic stirring for 2 h. After that, the mixture was introduced into a high-pressure reactor for a solvothermal reaction to prepare the final product CN-HMoP (Fig. 8c). Dai et al. [235] took the rod-shaped Bi-MOF named CAU-17 as a template and adopted a one-pot oil bath method to prepare the ZnIn2S4/Bi2S3 Z-scheme heterojunction photocatalyst. Specifically, a certain amount of CAU-17, InCl3·4H2O, ZnCl2 and thioacetamide (TAA) were thoroughly mixed in deionized water. Following the adjustment of the mixed solution's pH to 2.5, the reaction was carried out in an oil bath at 80 ℃ for 6 h. Ultimately, the yellow precipitate was harvested via centrifugation, subsequently washed with pure water and absolute ethanol, and dried to yield ZnIn2S4/Bi2S3 (Fig. 8d).

    Figure 8

    Figure 8.  (a) The synthetic process of the TCNx sample. Copied with permission [233]. Copyright 2024, Elsevier. (b) The illustration to prepare In2S3/MnIn2S4 catalyst. Copied with permission [234]. Copyright 2024, Elsevier. (c) Schematic showing the synthesis process of CN/HMoP synthesis. Copied with permission [127]. Copyright 2024, Elsevier. (d) Creative synthetic method of ZnIn2S4/Bi2S3 hybrids in this work. Copied with permission [235]. Copyright 2024, Wiley-VCH.

    In recent years, the continuous advancement of computer technology has given rise to a novel catalyst preparation strategy, namely the machine learning strategy. However, almost all the catalysts for producing H2O2 designed by the machine learning strategy are single-atom catalysts at present, and no heterojunction catalysts for producing H2O2 have been found to be designed yet. Herein, the process of designing catalysts for producing H2O2 by the machine learning strategy will be briefly introduced, aiming to provide researchers with new preparation ideas and explore the possibility of using the machine learning strategy to design heterojunction catalysts for producing H2O2 in the future. For example, Deng et al. [236] proposed an iterative machine learning (iML) approach, which can significantly diminish the required size of the training dataset. By incorporating spatial coordinate information as a feature, the objective of swiftly identifying the target catalyst with optimal catalytic activity from a vast array of single-atom catalysts can be realized. Concurrently, it has been established that RhN2O2A was an exemplary catalyst for photocatalytic H2O2 production, characterized by an exceedingly low overpotential of 0.013 V (Fig. 9a). Moreover, Zhang et al. [237] constructed five machine learning (ML) models, utilizing the adsorption free energies (ΔG(O*)) of 149 single-atom catalysts (SACs) and the limiting potentials (UL) of 31 SACs as the foundational data. Subsequently, descriptors capable of accurately describing SACs were obtained. Moreover, the two-electron oxygen reduction reaction (2e ORR) catalytic performances of 690 unknown SACs were well predicted, and four 2e ORR materials with relatively high selectivity and activity, namely Zn@Pc-N3C1, Au@Pd-N4, Au@Pd-N1C3 and Au@Py-N3C1, were screened out. Finally, the ULs of these SACs were verified through DFT calculations, and all were found to exceed the standard values, thereby demonstrating the efficacy of the ML models (Fig. 9b).

    Figure 9

    Figure 9.  (a) Scheme of iML method. Copied with permission [236]. Copyright 2022, Elsevier. (b) Selection methodology for SACs exhibiting elevated selectivity and catalytic activity. Copied with permission [237]. Copyright 2023, Elsevier.

    Characterization techniques on Z-scheme and S-scheme heterojunction catalysts for H2O2 production, such as X-ray diffraction (XRD), SEM, TEM, UV visible diffuse reflectance spectroscopy (UV–vis DRS), photoluminescence spectrum (PL), time-resolved photoluminescence spectrum (TRPL), electron paramagnetic resonance (EPR), a variety of photoelectrochemical testings, in-situ characterization techniques, density function theory (DFT) calculations and so on, especially, EPR, a variety of photoelectrochemical testings, in-situ characterization techniques and DFT calculations, provide significant guidance and basis for studying the charge transfer pathways and revealing the mechanisms of catalytic reactions. These characterizations also offer scientific research directions for the design of efficient Z-scheme and S-scheme heterojunction catalysts and the optimization of their performance. In this context, these characterization techniques are presented and discussed in detail.

    In the field of photocatalytic H2O2 production, EPR (usually also known as ESR) testing, a technique that uses an electron paramagnetic resonance spectrometer to detect unpaired electrons in materials, is frequently employed to identify the types of free radicals and their relative quantities. This helps researchers infer the charge-transfer mechanisms and catalytic mechanisms of Z-scheme or S-scheme heterojunction composites during the catalytic process.

    For example, Li et al. [128] utilized ESR experiments to characterize the primary reactive species involved in the photocatalytic process of 0.5S-pCN/WO2.72 S-scheme composite. As shown in Figs. 10a-c, under dark conditions, no characteristic peaks corresponding to DMPO-O2 and DMPO-OH were detected. However, upon light irradiation, distinct peaks for DMPO-O2 and DMPO-OH emerged, thereby confirming that O2 and OH were the predominant reactive species. In addition, the main reason for the insignificant change in the TEMPO-h+ signal was that some h+ reacted with H2O molecules and hydroxyl ions on the material surface to form OH, thus enhancing the DMPO-OH signal. Since only S-scheme can retain a high conduction band and a deep valence band potential, the results of the ESR test helped to confirm the inference that the 0.5S-pCN/WO2.72 composite formed an S-scheme heterojunction (Fig. 10d). Moreover, Liu et al. [122] conducted a series of EPR tests on the ZIS-Z/OCN Z-scheme heterojunction composite to explore the mechanism of photocatalytic H2O2 production. As shown in Fig. 10e, the absence of detectable H2O2 in the photoreaction system devoid of O2 and ZIS-Z/OCN underscores the indispensable roles of both O2 and ZIS-Z/OCN in the photocatalytic H2O2 production. Analogously, the introduction of superoxide dismutase (SOD) into the reaction system resulted in the nondetection of H2O2. In contrast, a small amount of H2O2 was monitored in the absence of IPA. The aforementioned results suggested that the presence of O2 was a fundamental prerequisite for the photocatalytic H2O2 production, thus inferring that H2O2 was not generated through a one-step two-electron method. Based on the EPR results in Figs. 10f-h, it can be deduced that both 1O2 and O2 were intermediate products in photocatalytic H2O2 production. Meanwhile, there was an energy transfer process during the photocatalytic H2O2 production, but 1O2 was mainly produced through the charge-transfer route. Consequently, the pathway of photocatalytic H2O2 production by ZIS-Z/OCN was inferred.

    Figure 10

    Figure 10.  ESR detection for DMPO-O2 (a), DMPO-OH (b) and TEMPO-h+ (c) under dark and visible-light irradiation, schematic illustrations of WO2.72 and 0.5S-pCN before and after contact, as well as the formation of IEF and band bending (d). Copied with permission [128]. Copyright 2023, Elsevier. The photocatalytic H2O2 production yield of 40ZIS-Z/OCN under various conditions (e), ESR spectra of DMPO-OH (f), DMPO-O2 (g) and TEMP-1O2 (h) for 40ZIS-Z/OCN under visible light irradiation. Copied with permission [122]. Copyright 2023, Elsevier. Transient photocurrent response (i), EIS analysis (j) and K-L curves (k) of ZnO, ZnIn2S4, and ZZS-20. Copied with permission [232]. Copyright 2023, Elsevier.

    There are numerous types of photoelectrochemical tests. Among them, electrochemical impedance spectroscopy (EIS) and transient photocurrent test (i-t) are frequently employed to assess and characterize the separation and migration efficiency of photogenerated e and h+ in Z-scheme or S-scheme heterojunction composites, thereby reflecting the photocatalytic activity of the materials. In addition, the slope of the curve in the linear sweep voltammetry (LSV) results can be calculated through the Koutecki-Levich (K-L) equation to estimate the number of electron transferred during the oxygen reduction reaction, thereby determining the reaction pathway during the photocatalytic H2O2 production.

    For example, Wu et al. [232] utilized i-t and EIS to evaluate the charge separation efficiency and transfer ability of the ZZS-20 S-scheme heterojunction composite. As shown in Figs. 10i and j, the photocurrent density of ZZS-20 was significantly elevated compared to that of ZnO and ZnIn2S4. This disparity attributed to the superior charge separation efficacy of the S-scheme heterojunction within the composite structure. Meanwhile, ZZS-20 exhibited the smallest arc radius, indicating that it had the strongest charge transfer ability. These findings collectively demonstrate that the construction of an S-scheme heterojunction was advantageous for enhancing the separation and transfer efficiency of photogenerated charges. Moreover, to elucidate the reaction pathways of different catalysts during the photocatalytic H2O2 production, rotational disk electrode tests were usually performed on materials to determine the number of electrons transferred (n) in ORR. Fig. 10k showed the slopes of LSV curves for ZnO, ZnIn2S4, and ZZS-20 calculated using K-L equation respectively. The calculated n value of ZZS-20 was 2.5, indicating a greater propensity for the two-electron transfer pathway in photocatalytic H2O2 production. Similarly, Jiang et al. [224] utilized the KL plot of SCN, TiO2, and SCN/T9 at −1.0 V vs. Ag/AgCl, which indicated that the n value of SCN/T9 was 2.12, thus inferring that the H2O2 production on SCN/T9 predominantly proceeded via the direct two-electron ORR pathway.

    In-situ characterization techniques have been pivotal in advancing the study of photocatalytic H2O2 production. These methodologies have not only enabled researchers to attain a profound understanding of the reaction mechanisms but also undergirded the development of more efficacious and sustainable protocols for H2O2 production. Balakrishnan et al. [238] employed in-situ fourier transform infrared spectroscopy (FTIR) to elucidate the mechanistic pathway of photocatalytic H2O2 production within the photocatalytic system supported by OPACN hydrogel (Figs. 11a-c). Their findings confirmed that the formation of H2O2 via photocatalysis on OPACN hydrogel proceeded via a two-step-two-electron reduction process. Zan et al. [239] capitalized on the semiconductor property that an increase in electron binding energy implies electron loss, and conversely, a decrease signifies electron gain. By employing in-situ X-ray photoelectron spectroscopy (XPS) technology, they observed the differences in electron binding energies of elements C, N and Bi under both dark and illuminated conditions to determine the direction of electron flow in the CNQDs/BiOBr composite material. The analysis of Figs. 11d-f indicated that an S-scheme heterojunction was successfully synthesized within the CNQDs/BiOBr composite, thereby conferring it with enhanced photocatalytic performance for the production of H2O2. Li et al. [240] utilized 5,5-dimethyl-1-pyrroline N-oxide (DMPO) as a radical scavenger to perform in-situ ESR characterization technique on various samples, thereby delving deeper into the mechanistic insights of the oxygen reduction reaction for photocatalytic H2O2 production using Au/F-TiO2 composite. As shown in Fig. 11g, the results indicated that superoxide radicals were formed by the initial combination of O2 with electrons and protons within the photocatalytic reaction medium. The generated HO2. radicals continued to react with one e and one H+, ultimately leading to the formation of H2O2. Consequently, H2O2 was progressively synthesized through a single-electron ORR on Au/F-TiO2 composite. Guo et al. [241] dissected the complex process of H2O2 production over an innovative 0D/2D carbon dots-modified CN nanosheet heterojunction (CDS10MCN) photocatalyst using in-situ diffuse infrared fourier transform spectroscopy (DRIFTS). As illustrated in Fig. 11h, the infrared spectra exhibited positive peaks corresponding to various intermediate species. Wang et al. [150] further verified the electron transfer mechanism between the PCN/MnS S-scheme heterojunction during the photocatalytic process by using in-situ Kelvin probe force microscopy (KPFM), which helped to deduce the generation mechanism of H2O2. As shown in the results of Figs. 11i-k, upon light irradiation, the surface potential of MnS exhibited a decrease, whereas the surface potential of PCN showed an increase. Therefore, under the action of the IEF, photoelectrons transfered from PCN to MnS, thereby augmenting the reduction capacity of the photogenerated electrons and consequently facilitating the photocatalytic H2O2 production.

    Figure 11

    Figure 11.  In-situ FTIR spectra of a series of OPACN composites during the photocatalytic H2O2 production (a) and an enlarged view of the FTIR peak (b, c). Copied with permission [238]. Copyright 2024, Elsevier. (d-f) In-situ XPS spectra of C, N and Bi in CNQDs/BiOBr-1.50%. Copied with permission [239]. Copyright 2023, Editorial Office of Acta Physico-Chimica Sinica. (g) In-situ ESR spectra of 0.1% Au/TiO2 composites and pure TiO2. Copied with permission [240], Copyright 2021, MDPI. (h) In-situ DRIFTS of CDs10MCN collected in the H2O2 photocatalytic processes in the O2 atmosphere. Copied with permission [241]. Copyright 2024, Wiley-VCH. In-situ KPFM potential diagram in dark (i), under light irradiation (j), and surface potential (k) of PCN(5)/MnS-D. Copied with permission [150]. Copyright 2023, Elsevier.

    DFT calculations can be utilized to clarify the impact of catalyst selection on the photocatalytic production rates of H2O2. This analysis also permits a deeper understanding of the relationship between the surface properties of the catalyst and the dynamics of the catalytic reaction, among other aspects. Zhang et al. [242] demonstrated through DFT calculations that the reduction in photocatalytic H2O2 decomposition may be attributed to the suppression of H2O2 adsorption by the photocatalyst. As shown in Fig. 12a, the TiO2 (101) surface exhibited the most pronounced adsorption affinity for H2O2. Conversely, as depicted in Fig. 12b, the maximum adsorption energy between H2O2 and MoSe2 was more positive due to lower surface reactivity, hence the absolute adsorption energy of H2O2 on core-shell biphase (1T-2H)-MoSe2/TiO2 nanoribbons (NRAs) was significantly lower than that on TiO2 NRAs. These findings indicated that the (1T-2H)-MoSe2 on TiO2 NRAs adsorbs less H2O2, thereby inhibiting the decomposition of H2O2. Guo et al. [241] employed DFT calculations to further elucidate the complex interplay between the electronic structures of CDs and CDs10MCN and its impact on reaction kinetics. As illustrated in Figs. 12c and d, the integration of CDs significantly enhanced the negative charge density of the CDs10MCN heterojunction, which in turn reduced the adsorption capacity for OOH*. The results demonstrated that the adsorption efficiency of H2O2 can be enhanced via 2e ORR pathway. Li et al. [243] elucidated the influence of the local coordination environment of zinc sites on the catalytic activity for oxygen reduction to H2O2 through theoretical calculations. As depicted in Figs. 12e-k, they identified a robust electronic interaction between Zn and O atoms, with the adsorption free energy of O2 on Zn-N3O being lower compared to Zn-N4. Furthermore, Zn-N3O facilitated the accumulation of electrons in *OOH. In summary, adsorbed oxygen was more readily activated to *OOH on Zn-N3O compared to Zn-N4 and subsequently converted into H2O2.

    Figure 12

    Figure 12.  Optimal configurations of H2O2 adsorption on TiO2 (101) (a) and 2H-MoSe2 (103) planes (b). Copied with permission [242]. Copyright 2022, Elsevier. Optimized unit cells of states for CDs10MCN (c) and free energy diagram of CDs10MCN for photocatalytic H2O2 production (d). Copied with permission [241]. Copyright 2024, Wiley-VCH. (e-k) The theoretical calculation results regarding the Zn-N4 and Zn-N3O models. Copied with permission [243]. Copyright 2024, Wiley-VCH.

    Nowadays, Z-scheme or S-scheme heterojunction catalysts for photocatalytic and photoelectrocatalytic H2O2 production are quite common. However, with the in-depth development of research in the catalytic field, researchers no longer pursue the design of catalysts with only single-benefit. Therefore, an increasing number of multi-functional and multi-benefit catalysts have emerged, which are not limited to merely achieving the single function of photocatalytic or photoelectrocatalytic H2O2 production. For instance, the coupling of photocatalytic or photoelectrocatalytic H2O2 production with the degradation of pollutants, as well as heterojunction catalysts that can respectively achieve functions such as efficient H2O2 production, H2 production, and nitrogen fixation. Here, we will outline some Z-scheme or S-scheme heterojunction catalysts for photocatalytic and photoelectrocatalytic H2O2 production with multi-functional applications, including the utilization of in-situ photocatalytic and photoelectrocatalytic H2O2 production for self-Fenton degradation.

    5.1.1   Z-scheme heterojunction catalysts for photocatalytic H2O2 production

    Specifically, Balakrishnan et al. [238] employed pectin as the foundational material and utilized distinct functionalized graphitic carbon nitrates (GCN), amino-rich GCN (ACN), oxidized GCN (OCN), along with phosphorylated GCN (PCN) through a wet impregnation method to effectively construct 3D OPACN hydrogel dual Z-scheme heterojunction catalyst. It is capable of facilitating the efficient degradation of tetracycline (TC), the photocatalytic H2O2 production and the fixation of N2 under photocatalysis. Fig. 13a depicted the photoluminescence (PL) spectrum of GCN-based catalysts. Analysis of the spectrum revealed that the PL intensity of the catalysts was attenuated post-modification. Specifically, the OPACN hydrogel demonstrated the minimal PL intensity among the catalysts evaluated, suggesting that the incorporation of pectin with the OPACN has enhanced the separation efficiency of photogenerated charge carriers. The efficiency of photogenerated charge carrier separation for the GCN-based catalysts were further elucidated through electron impedance spectroscopy (EIS). As illustrated in Fig. 13b, the OPACN hydrogel exhibited the minimum arc radius, indicating that, within the spectrum of catalysts under assessment, it possessed the most outstanding charge-carrier separation efficiency. The Mott-Schottky (M-S) curves of ACN, OCN and PCN composite materials were shown in Fig. 13c. As can be observed from Fig. 13c, the M-S curves of all three samples exhibited positive slopes. This phenomenon served as an indication that they are all n-type semiconductors. Moreover, it was possible to estimate that the conduction band potentials of OCN, ACN and PCN were −0.907, −0.697 and −0.217 eV respectively. Additionally, the authors utilized the iodometric method to measure the content of photocatalytically H2O2 production. It was ultimately concluded that the OPACN hydrogel could photocatalytically generate 1204 µmol/L of H2O2 within a 60-min period (Fig. 13d). After 6 times of experiment (Fig. 13e), the production of H2O2 only decreased by 56 µmol/L. Ultimately, Fig. 13f showed the reaction mechanism of OPACN hydrogel photocatalytically degrading tetracycline, fixing nitrogen, and producing H2O2 under visible light. OPACN hydrogel suppressed the recombination of carriers by constructing dual Z-scheme heterojunction, ensuring higher carrier separation efficiency in the system. At the same time, by changing the band gap, it achieved better visible light utilization. In addition, in OPACN hydrogel, the effective interaction between pectin and OPACN through hydrogen bonds led to a higher electron transfer rate in the system and promoted the generation of superoxide radicals, making the OPACN system had strong photocatalytic activity and reusability.

    Figure 13

    Figure 13.  Photoluminescence spectra (a), EIS Nyquist plot (b), Mott-schottky plot (c) and the yield of H2O2 (d) for the catalysts. (e) The results of repeatability experiments on the production of H2O2 by the OPACN hydrogel. (f) The photocatalytic mechanism illustration of OPACN hydrogel. Copied with permission [238]. Copyright 2024, Elsevier.

    Ji et al. [244] have engineered an innovative Z-scheme heterojunction photocatalyst consisting of cobalt-iron oxide and perylene diimide organic supramolecule (CoFeO/PDIsm). Within the photocatalytic-self-Fenton system, CoFeO/PDIsm utilized photogenerated h+ for the mineralization of organic pollutants, while the photogenerated e efficiently produced in-situ H2O2. As a result, CoFeO/PDIsm exhibited an excellent in-situ photocatalytic H2O2 production rate of 281.7 µmol g-1 h-1 (Fig. 14a) in the pollutant solution, and the corresponding total organic carbon (TOC) removal rate of ciprofloxacin (CIP) was 63.7% (Fig. 14b), which far exceeded that of existing photocatalysts. In addition, CoFeO/PDIsm exhibited stable and remarkable degradation capabilities for various organic pollutants (Fig. 14c). After multiple cycle experiments, CoFeO/PDIsm still exhibited stable photocatalytic performance. As shown in Fig. 14d, under certain experimental conditions, CoFeO/PDIsm exhibited the best degradation rate constant value (K) of 0.0445 min-1 for CIP in the photocatalytic self-Fenton system under an air atmosphere. From this, it can be judged that the excellent degradation efficiency in the CoFeO/PDIsm system mainly comes from the in-situ H2O2 activated by Fe(Ⅱ). The findings presented in Fig. 14e indicated that, within the photocatalytic self-Fenton system, Fe3+ was efficiently reduced to Fe2+. This reduction further enhanced the utilization efficiency of H2O2, consequently leading to the generation of a substantial quantity of OH radicals. Specifically, in the photocatalytic self-Fenton system, CoFeO/PDIsm produced OH radicals at a high concentration, with a generation rate of 144.5 µmol L-1 h-1. This rate was 5.7-fold higher than that in the Fenton system, as shown in Fig. 14f. Moreover, CoFeO/PDIsm exhibited an outstanding OH conversion rate of 50%, significantly surpassing that achieved in the Fenton reaction (Fig. 14g). Thus, it was inferred that the high mineralization ability of CoFeO/PDIsm mainly stemmed from the rapid conversion of in-situ photocatalytic H2O2 production to OH in the photocatalytic self-Fenton system. Ultimately, as depicted in Fig. 14h, the potential mechanism of CIP degradation by CoFeO/PDIsm within the photocatalytic self-Fenton system was proposed. Upon excitation of CoFeO/PDIsm by visible light, the holes in the valence band of PDIsm underwent recombination with the electrons in the conduction band of CoFeO. Subsequently, electrons with enhanced reducing capacity were retained in the conduction band of PDIsm, which could be utilized for the in-situ photocatalytic H2O2 production. In the meantime, the h+ with higher oxidation ability in the VB of CoFeO combined with OH, which was utilized for the mineralization of CIP.

    Figure 14

    Figure 14.  (a) H2O2 production of the samples. (b) The comparison of TOC removal rate. (c) The TOC removal rate of various organic pollutants over CoFeO/PDIsm. (d) The rate constants of the reaction for the degradation of CIP under diverse systems. (e) The concentration of Fe2+ under diverse systems. (f) The concentration of radical OH under diverse systems (f). (g) OH transformation rate under diverse systems. (h) The postulated mechanism for the degradation reaction induced by CoFeO/PDIsm. Copied with permission [244]. Copyright 2023, Elsevier.
    5.1.2   Z-scheme heterojunction catalysts for photoelectrocatalytic H2O2 production

    Feng et al. [245] synthesized a yolk-shell structured Z-scheme heterojunction composite, designated as YS-CuCo2S4@Cu2O-NR, with an octahedral copper oxide core and a tubular CuCo2S4 shell, by adjusting the composition and morphology through a hydrothermal method. YS-CuCo2S4@Cu2O-NR possesses upper-level photocatalytic activity, which can be synergized by spatial confinement effects to selective oxidation of benzyl alcohol (BA) to generate high value-added benzaldehyde (BAD)/degradation of multiple pollutants using Fenton-like reaction, with 80% BA conversion, 99% selectivity (Fig. 15a), 12 mmol L-1 g-1 yield to H2O2, and >90% degradation efficiency for multiple pollutants. Furthermore, the time-dependent conversion of BA as depicted in the HPLC chromatogram presented in Fig. 15b did not exhibit over-oxidation products. As illustrated in Fig. 15c, for the selective oxidation of BA by YS-CuCo2S4@Cu2O-NR, the conversion and selectivity reached 79.4% and 99.2% respectively at 2.5 h. Additionally, the recyclability of YS-CuCo2S4@Cu2O-NR was investigated through conducting four cyclic photocatalytic processes for BAD production, with each cycle lasting 2.5 h. After four cycles, it retained approximately 88% of its initial activity (Fig. 15d). To delve into the potential oxidation processes of BA, trapping experiments and ESR tests were carried out. As demonstrated in Fig. 15e, the results of the trapping experiments revealed that photogenerated holes and O2 radicals were of vital importance in the photocatalytic oxidation of BA. The DMPO-O2 and DMPO-C radicals were detected, and the characteristic peak signals augmented as time elapsed, suggesting that the C radical also played a pivotal role in the photocatalytic oxidation of BA (Fig. 15f).

    Figure 15

    Figure 15.  The selective photocatalytic oxidation of benzaldehyde over various samples (a), the HPLC chromatogram depicting benzaldehyde and the resultant benzoin dimethyl acetal generated within YS-CuCo2S4@Cu2O-NR (b), time-dominated selective oxidation of benzaldehyde (c), the recyclability of YS-CuCo2S4@Cu2O-NR during the oxidation of benzaldehyde accompanied by the production of H2O2 (d), EPR signals recorded for YS-CuCo2S4@ Cu2O-NR under light illumination (e), DMPO-C/O2− in benzaldehyde with YS-CuCo2S4@Cu2O-NR (f). Copied with permission [245]. Copyright 2024, Elsevier.

    Sun et al. [246] prepared a 3D hollow-sphere NiFe2O4@ZnFe2O4 (NF@ZF) Z-scheme heterojunction catalyst via the hydrothermal method. Subsequently, an asymmetric-wettability Janus photoelectrode was constructed by loading hydrophobic polytetrafluoroethylene (PTFE) on one face of three-dimensional porous graphite felt (GF) and NF@ZF on the opposite face. This Janus photoelectrode was characterized by a hydrophobic oxygen-storage layer and a hydrophilic catalyst layer. As illustrated in Fig. 16a, the synthesized Janus photocathode demonstrated remarkable photoelectric characteristics, featuring a low onset potential and a swift increase in current density. Moreover, the impact of electrode potential on the H2O2 yield was explored by quantitatively measuring the H2O2 concentration subsequent to 1 h of photoelectrochemical reduction of O2. As presented in Fig. 16b, the H2O2 yield of the Janus photocathode with asymmetric wettability achieved a maximum of 255.9 mg L-1 h-1 at −0.7 V versus Ag/AgCl, accompanied by a Faraday efficiency (FEs) of 72% (Fig. 16c). To assess the stability of the Janus electrode, a long-term photoelectrochemical oxygen reduction reaction was conducted. After 9 h of testing, there was scarcely any alteration in the current (Fig. 16d). The H2O2 yield exhibited an approximately linear correlation with the catalytic duration, and the Faraday efficiency remained steadily within the range of 60%−80% (Fig. 16e). Additionally, the oxidation of arsenite (As(Ⅲ)) to arsenate (As(Ⅴ)) and the transformation/degradation capability of roxarsone (ROX) in the Janus‖Fe in-situ Fenton system were investigated. The results suggested that a potential of −0.7 V was adequate for the oxidation process, and the optimal pH value of the system was 3 (Fig. 16f). The degradation efficiency of ROX reached nearly 100% within 120 min, while 20 ppm of As(Ⅲ) was almost entirely oxidized to As(Ⅴ) within 45 min (Fig. 16g). Finally, the oxidation mechanism of As(Ⅲ) in the Janus‖Fe in-situ Fenton system was shown in Fig. 16h.

    Figure 16

    Figure 16.  LSV curves of different electrodes (a), the influence of different applied potentials on the production of H2O2 (b), H2O2 yields and FEs of different electrodes (c), stability experiment for the H2O2 production via PEC (d), H2O2 yield and FEs of the Janus electrode (e), the influence of different conditions on As(Ⅲ) oxidation and ROX degradation (f), both the conversion efficiency of As(Ⅲ) and the degradation efficiency of ROX over different systems (g), schematic illustration of the mechanism underlying the photoelectrochemical process (h). Copied with permission [246]. Copyright 2024, Elsevier.
    5.2.1   S-scheme heterojunction catalysts for photocatalytic H2O2 production

    Jing et al. [247] prepared an S-scheme heterojunction composite (MCN@CdS) by loading CdS nanoparticles onto square-tubular g-C3N4 (MCN) via the solvothermal method. Under visible-light irradiation, MCN@CdS exhibited excellent dual photocatalytic properties. It can not only photocatalytically reduce oxygen to produce H2O2, but also had a strong ability to photocatalytically degrade antibiotics. As shown in Figs. 17a and b, the H2O2 production rate of MCN@CdS can reach 0.95 mmol/L within 90 min, and the degradation rate of TC was as high as 99.5% within 120 min. After five cyclic experiments, each lasting for 90 min, the photocatalytic production of H2O2 by MCN@CdS decreased by only 0.1 mmol/L, indicating that MCN@CdS had a certain degree of stability. The larger rate constant (k value) of MCN@CdS also reflected its better photocatalytic performance (Fig. 17c). In addition, MCN@CdS exhibited the lowest emission intensity in Fig. 17d, indicating that MCN@CdS had the highest separation efficiency of photo-carriers. As shown by the results of time-resolved PL spectroscopy (TRPL) in Fig. 17e, the composite exhibited a longer carrier lifetime. Moreover, the superior photoelectric properties of MCN@CdS were verified in Figs. 17f and g. The results indicated that MCN@CdS was more conducive to charge transfer and had a higher photocurrent intensity under visible light. Finally, the mechanism of the coupling of H2O2 production and TC degradation by photocatalysis over MCN@CdS S-scheme heterojunction catalyst was presented in Fig. 17h.

    Figure 17

    Figure 17.  Photocatalytic H2O2 production by the samples (a), degradation performance (b) and the corresponding degradation kinetics (c) of TC by the samples, PL (d), TRPL (e) and EIS (f) spectra, along with photocurrent response (g) of different samples, mechanism diagram of the coupling of H2O2 production and TC degradation by photocatalysis of MCN@CdS S-scheme heterojunction composite (h). Copied with permission [247]. Copyright 2024, Elsevier.

    Du et al. [248] constructed a photo-self-Fenton system based on resorcinol-formaldehyde resin/MIL-88A(Fe) (RF/88A) S-scheme composite. This system facilitates the in-situ photocatalytic H2O2 production when exposed to visible-light irradiation, which is then utilized for the degradation of TC and sterilization purposes. As shown in Fig. 18a, within 180 min, RF/88A can achieve a degradation efficiency of over 80% for TC upon visible-light irradiation. In addition, within the tests of antibacterial activity, 35%-RF/88A nearly eliminated 100% of both Escherichia coli (E. coli) and Staphylococcus aureus (S. aureus) in 60 min upon visible-light irradiation (Figs. 18b-e). After five consecutive repeated experiments, each lasting for 3 h, the efficiency of photocatalytic self-Fenton degradation of TC by 35%-RF/88A decreased by <10%, which proved that 35%-RF/88A exhibited substantial photocatalytic stability. Moreover, Sun et al. [249] constructed a photocatalysis-self-Fenton system built upon a 2D/2D Bi2Fe4O9/ZnIn2S4 (BFO/ZIS) S-scheme catalyst, which was capable of achieving highly efficient degradation of antibiotic pollutants. As shown in Fig. 18f, under pure water conditions without additional oxygen supply and sacrificial agents, in 2 h, the H2O2 production efficiency via photocatalysis of BFO/ZIS approached 30 µmol/L. This is because after the formation of the heterostructure, the decrease in the photocatalytic H2O2 concentration in BFO/ZIS may be ascribed to the Fenton reaction that occurs between H2O2 and iron species throughout the process of in-situ photocatalytic H2O2 production, leading to the activation of H2O2 to generate OH radicals. The BFO/ZIS composite with the optimal performance achieved an excellent degradation efficiency of 88.8% for TC after 2 h of visible-light illumination (Fig. 18g). After four consecutive cyclic tests, each lasting for 2.5 h, BFO/ZIS still showed remarkable photocatalytic self-Fenton degradation efficiency for TC. Finally, the potential mechanism of antibiotic degradation by the photocatalysis-self-Fenton system was proposed in Fig. 18h. Specifically, in BFO/ZIS, ZIS promoted the in-situ production of H2O2 from deionized water through photocatalysis, while BFO provided Fe2+ as a Fenton reagent. The synergistic effect between them constituted a self-sufficient photocatalysis-self-Fenton system, which was employed for the degradation of antibiotics.

    Figure 18

    Figure 18.  TC degradation efficiency of diverse samples (a), the cell density of E. coli (b) and S. aureus (c) for diverse samples, the antibacterial performances of diverse samples for E. coli (d) and S. aureus (e). Copied with permission [248]. Copyright 2024, Elsevier. Concentration of H2O2 production curves for different samples (f), the TC degradation performance upon diverse conditions (g), possible mechanism for TC degradation of BFO/ZIS (h). Copied with permission [249]. Copyright 2024, Elsevier.
    5.2.2   S-scheme heterojunction catalysts for photoelectrocatalytic H2O2 production

    Recently, Chen et al. [234] employed the ion exchange strategy to create a hollow tubular In2S3/MnIn2S4 S-scheme composite. Capable of triggering the ORR on the hydrophobic In2S3/MnIn2S4/PVDF/NF photo-anode and the ClOR on the hydrophilic In2S3/MnIn2S4/CP photo-cathode, the catalyst led to the build-up of practical concentrations of H2O2 and HClO, attaining levels as high as 2108 µmol/L and 28.5 mg/L respectively. Through the utilization of a dual-electrode co-photoelectrocatalytic system involving O2 and Cl, which replaced the oxygen evolution reaction (OER) with ClOR, H2O2 and HClO were generated simultaneously at a lower voltage in an H-type cell (Fig. 19a). In a single-chamber cell, the electrodes can trigger the activation of H2O2 and HClO, which then act in synergy to yield singlet oxygen (1O2) and numerous hydroxyl radicals (OH). This combined action effectively leads to the breakdown of organic pollutants. Fig. 19b illustrated the linear sweep voltammetry (LSV) curves for the various systems. Compared to the PEC ORR||OER system, the PEC ORR||ClOR electrolysis demanded a relatively reduced potential, which effectively illustrated the energy-conservation benefit of supplanting OER with ClOR. As depicted in Fig. 19c, under the condition of 1.2 V Ag/AgCl, the accumulated concentrations of H2O2 and HClO within 2 h respectively reached 1.49 mmol/L and 17.3 mg/L. Fig. 19d demonstrated that the electrolytic cell coupled with both reactions exhibited good yields for both H2O2 and HClO, and the photolytic stability could be maintained for up to 8 h. Additionally, Fig. 19e presented the results of MB degradation for each system, with the H2O2-HClO coupled system exhibiting the strongest degradative capability. Combining the analysis of various characterization results, the authors have proposed the mechanism for the improved Fenton-like degradation by the H2O2-HClO coupled system when exposed to visible light (Fig. 19f). This work provides a synergistic strategy for the concurrent generation of H2O2 and HClO acid, offering an efficient solution for pollutant degradation in the realms of environmental and energy engineering. Finally, as depicted in Fig. 19g, the complete mechanism of In2S3/MnIn2S4 photoelectrocatalytic system for degrading organic pollutants through the production of H2O2/HClO was presented. In addition, Table S1 (Supporting information) summarized representative Z-scheme and S-scheme heterojunction catalysts for the production of H2O2 in the fields of photocatalysis [250-260].

    Figure 19

    Figure 19.  H2O2 production at the anode (a), LSV plots of In2S3/MnIn2S4/PVDF/NF||In2S3/MnIn2S4/CP (b), different potentials for the production of H2O2 and HClO (c), recyclability and stability of ORR||ClOR system (d), degradation efficiency of MB in different systems (e), mechanism of OH activation (f), mechanism diagram of organic pollutant degradation by H2O2/HClO production in In2S3/MnIn2S4 photoelectrocatalytic system (g). Copied with permission [234]. Copyright 2024, Elsevier.

    This review presents an exhaustive review of the mechanisms underpinning the photocatalytic and photoelectrocatalytic H2O2 production, utilizing Z-scheme and S-scheme heterojunction catalysts. Additionally, it discusses several research findings pertaining to the photocatalytic and photoelectrocatalytic H2O2 production. In particular, the resolution of H2O2 decomposition poses a significant challenge, and the development of effective strategies for H2O2 extraction and photocatalyst recovery is of paramount importance. Despite the notable advancements in Z-scheme and S-scheme catalysts for photocatalytic and photoelectrocatalytic H2O2 production in recent years, the quest for catalysts that exhibit high activity, selectivity, and stability continues to be a pivotal focus within the domains of photocatalytic and photoelectrocatalytic H2O2 production.

    The catalytic systems and their operational environments are also undergoing continuous evolution [261-265]. The fabrication of bifunctional catalysts capable of mediating both 2e ORR and 2e WOR presents expanded opportunities for the efficient photocatalytic and photoelectrocatalytic H2O2 production and the effective partitioning and exploitation of photogenerated electron-hole pairs. To augment the separation efficiency of these photogenerated charge carriers, the utilization of sacrificial agents is a prevalent strategy. However, these agents pose a significant environmental pollution risk. Consequently, there is a pressing need to prioritize the development of catalytic systems that can operate without the reliance on sacrificial agents, thereby mitigating their environmental impact. In addition to the use of pure water systems, the production of H2O2 from seawater is beginning to receive attention. The implementation of a seawater environment further alleviates the constraints on the preparatory conditions necessary for H2O2 production. H2O2 can be synthesized under non-extreme, nature-simulating conditions, which is undoubtedly a significant milestone in harnessing energy from natural sources and represents an exciting advancement in the field. However, it must be acknowledged that the current system is confined to laboratory-scale operations, and further efforts are imperative to translate this technology into practical industrial applications.

    In the realm of photocatalytic and photoelectrocatalytic H2O2 production, there exist several avenues of thought that warrant in-depth discussion and research [266-270]. Fig. 20 mainly summarizes the potential development directions of Z-scheme and S-scheme heterojunction materials in the field of photocatalytic and photoelectrocatalytic H2O2 production. First and foremost, enhancing the oxidation resistance and photocorrosion resistance of both the ion-exchange membrane and the catalyst is imperative for ensuring the long-term stability within an environment rich in high-concentration H2O2 [271-275]. This enhancement is pivotal in substantially improving the cyclic stability of the integral reaction system. Secondly, enhancing the mass transfer efficiency within the device is achievable through a meticulous redesign of the flow cell's channel dimensions and their spatial configuration. In parallel, the strategic placement of spacer groups at the interfacial region can further contribute to a significant improvement in mass transfer efficiency [276-280]. Thirdly, the decomposition of H2O2 is a critical factor that must be addressed. By inhibiting the decomposition of H2O2 during the production process and optimizing the collection methodology, the overall production efficiency of the system can be significantly enhanced [281-285]. Fourthly, beyond the photocatalytic and photoelectrocatalytic H2O2 production within a standalone system, integrating the H2O2 production reaction system with other processes, such as wastewater treatment, hydrogen generation, and organic synthesis, can construct a catalytic system with enhanced economic value. This approach aims to achieve a multifunctional and highly efficient H2O2 production reaction [286-290]. Fifthly, it is essential to explore superior pathways for the catalytic reaction aimed at enhancing the high activity and high selectivity in the photocatalytic and photoelectrocatalytic synthesis of H2O2 [291-294]. Sixthly, efforts should be directed towards immobilizing the catalyst on a membrane structure to facilitate the efficient production of H2O2. Concurrently, this approach enhances the catalyst's recyclability, thereby realizing a catalytic system that is both sustainable and facile in terms of recovery and reuse [295-297]. Seventhly, promote the process of implementing photocatalytic H2O2 production from the laboratory stage to the industrialization stage. Ultimately, it is anticipated that in the imminent future, the possibility of using machine learning strategies to design heterojunction catalysts for the photocatalytic and photoelectrocatalytic H2O2 production can be explored, so as to enhance the scientificity and efficiency of catalyst preparation [298-301].

    Figure 20

    Figure 20.  A schematic diagram of the research directions and development prospects of photocatalytic and photoelectrocatalytic H2O2 production.

    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.

    Xibao Li: Writing – review & editing, Writing – original draft, Visualization, Validation, Supervision, Software, Resources, Project administration, Methodology, Investigation, Funding acquisition, Formal analysis, Data curation. Yiyang Wan: Writing – original draft, Visualization, Validation, Software, Resources, Methodology, Investigation, Formal analysis, Data curation. Fang Deng: Writing – original draft, Software, Resources, Methodology, Investigation, Formal analysis, Data curation. Yingtang Zhou: Writing – original draft, Software, Resources, Methodology, Investigation, Formal analysis, Data curation. Pinghua Chen: Writing – original draft, Validation, Software, Resources, Methodology, Investigation. Fan Dong: Writing – review & editing, Software, Resources, Methodology, Investigation, Funding acquisition. Jizhou Jiang: Writing – review & editing, Visualization, Validation, Supervision, Software, Resources, Project administration, Methodology, Investigation, Funding acquisition, Formal analysis, Data curation, Conceptualization.

    The authors gratefully acknowledge the financial support provided by the National Natural Science Foundation of China (Nos. 22262024, 52470078, 62004143), Jiangxi Province Academic and Technical Leader of Major Disciplines (No. 20232BCJ22008), the Key Project of Natural Science Foundation of Jiangxi Province (Nos. 20232ACB204007), Double Thousand Talent Plan of Jiangxi Province, the Natural Science Foundation of Jiangxi Province (No. 2022ACB203014), the Key R&D Program of Hubei Province (No. 2022BAA084), and the Innovation Project of Engineering Research Center of Phosphorus Resources Development and Utilization of Ministry of Education (No. LCX202404).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2025.111418.


    1. [1]

      K. Kamata, K. Yonehara, Y. Sumida, et al., Science 300 (2003) 964–966. doi: 10.1126/science.1083176

    2. [2]

      J. Wang, X. Cao, X. Cui, et al., Adv. Mater. 36 (2024) 2311151. doi: 10.1002/adma.202311151

    3. [3]

      S.B. Zhu, G.X. Zhang, Y.L. Bao, et al., Rev. Adv. Mater. Sci. 62 (2023) 20220293. doi: 10.1515/rams-2022-0293

    4. [4]

      Y. Zhang, J. Huang, Z.X. Dong, et al., Carbon Lett. 33 (2023) 77–87. doi: 10.1007/s42823-022-00404-z

    5. [5]

      J.S. Shen, Q. Liu, Q.Q. Sun, et al., J. Ind. Eng. Chem. 118 (2023) 170–180. doi: 10.1016/j.jiec.2022.11.002

    6. [6]

      M. Tayyab, U. Kulsoom, Y. Liu, et al., Int. J. Hydrogen Ener. 51 (2024) 1400–1413. doi: 10.1016/j.ijhydene.2023.09.199

    7. [7]

      Z. Tang, P. Zhao, H. Wang, et al., Chem. Rev. 121 (2021) 1981–2019. doi: 10.1021/acs.chemrev.0c00977

    8. [8]

      R.A. He, D.F. Xu, X. Li, J. Mater. Sci. Technol. 138 (2023) 256–258. doi: 10.1016/j.jmst.2022.09.002

    9. [9]

      Y.C. Guan, Z.X. Ren, Y. Lang, et al., Rev. Adv. Mater. Sci. 62 (2023) 20220280. doi: 10.1515/rams-2022-0280

    10. [10]

      D. Ozturk, M. Gülcan, J. Ind. Eng. Chem. 122 (2023) 251–263. doi: 10.1016/j.jiec.2023.02.026

    11. [11]

      S.Z. Li, T.Y. Wang, Q. Li, Sci. China Chem. 66 (2023) 3398–3414. doi: 10.1007/s11426-023-1760-x

    12. [12]

      H. Li, B.C. Zhu, B. Cheng, et al., J. Mater. Sci. Technol. 161 (2023) 192–200. doi: 10.1016/j.jmst.2023.03.039

    13. [13]

      Y. Li, P.Z. Xie, Y.Z. Hui, et al., Int. J. Hydrogen Energ. 58 (2024) 596–604. doi: 10.1016/j.ijhydene.2024.01.211

    14. [14]

      X. Wang, C.Z. Chen, Q.S. Wang, K. Dai, Chin. J. Struct. Chem. 43 (2024) 100473.

    15. [15]

      W. Iqbal, R.N. Liu, Y.P. Mao, et al., Mater. Today Chem. 41 (2024) 102315. doi: 10.1016/j.mtchem.2024.102315

    16. [16]

      J. Xu, X. Zheng, Z. Feng, et al., Nat. Sustain. 4 (2021) 233–241.

    17. [17]

      C.W. Zhao, W. Li, J.S. Hu, et al., Environ. Res. 261 (2024) 119775. doi: 10.1016/j.envres.2024.119775

    18. [18]

      Z. Li, W. Zhou, M. Yu, et al., J. Liaocheng Univ. (Nat. Sci. Ed.) 37 (2024) 50–61.

    19. [19]

      Y.Y. Cui, J.F. Zhang, H.L. Chu, L.X. Sun, K. Dai, Acta Phys. Chim. Sin. 40 (2024) 2405016. doi: 10.3866/pku.whxb202405016

    20. [20]

      S.J. Sun, J. Zhang, C.D. Sheng, et al., J. Hazard. Mater. 440 (2022) 129788. doi: 10.1016/j.jhazmat.2022.129788

    21. [21]

      C.G. Chen, J.F. Zhang, H.L. Chu, et al., Chin. J. Catal. 63 (2024) 81–108. doi: 10.1016/S1872-2067(24)60072-0

    22. [22]

      X.L. Zhang, C.Y. Wu, Z.H. Wang, et al., Environ. Res. 263 (2024) 120020. doi: 10.1016/j.envres.2024.120020

    23. [23]

      Y. Wang, Y.F. Wang, W. Hu, et al., Chem. Commun. 60 (2024) 7331–7334. doi: 10.1039/d4cc01576b

    24. [24]

      J.X. Xie, L.J. Zhong, X. Yang, et al., Chin. Chem. Lett. 35 (2024) 108472. doi: 10.1016/j.cclet.2023.108472

    25. [25]

      X. Luo, R.Y. Zhu, L. Zhao, et al., Environ. Res. 251 (2024) 118644. doi: 10.1016/j.envres.2024.118644

    26. [26]

      Y.Q. Bian, H.W. He, G. Dawson, J.F. Zhang, K. Dai, Sci. China Mater. 67 (2024) 514–523. doi: 10.1007/s40843-023-2725-y

    27. [27]

      T.C. Miao, X.M. Lv, F.S. Chen, et al., J. Colloid Interf. Sci. 663 (2024) 413–420. doi: 10.1016/j.jcis.2024.02.175

    28. [28]

      Y. Gao, W.H. Zhu, Y.Q. Li, et al., J. Environ. Manage. 304 (2022) 114315. doi: 10.1016/j.jenvman.2021.114315

    29. [29]

      J.X. Zhang, P.F. Tian, A.H. Xu, et al., Chin. Chem. Lett. 34 (2023) 108446. doi: 10.1016/j.cclet.2023.108446

    30. [30]

      P. Ruban, L.J. Reddy, M. Rajalakshmi, et al., Rev. Adv. Mater. Sci. 62 (2023) 20220301. doi: 10.1515/rams-2022-0301

    31. [31]

      Z.Q. Zheng, F. Han, B. Xing, et al., J. Colloid. Interf. Sci. 624 (2022) 460–470. doi: 10.1016/j.jcis.2022.05.161

    32. [32]

      B. Zhang, S. Wang, C.T. Qiu, et al., Appl. Surf. Sci. 573 (2022) 151506. doi: 10.1016/j.apsusc.2021.151506

    33. [33]

      J. Ma, X.X. Peng, Z.X. Zhou, Y.F. Shen, Y.J. Zhang, Chin. Chem. Lett. 34 (2023) 108784. doi: 10.1016/j.cclet.2023.108784

    34. [34]

      C.C. Liu, H.L. Tong, P.F. Wang, et al., Chem. Eng. J. 476 (2023) 146573. doi: 10.1016/j.cej.2023.146573

    35. [35]

      Y.X. Xie, G. Zhang, Mater. Lett. 373 (2024) 137147. doi: 10.1016/j.matlet.2024.137147

    36. [36]

      X.D. Wang, Y.N. Liu, Z.B. Liu, et al., Chin. Chem. Lett. 35 (2024) 108926. doi: 10.1016/j.cclet.2023.108926

    37. [37]

      L. Yu, Y. Wang, J. Zou, et al., J. Liaocheng Univ. (Nat. Sci. Ed.) 36 (2023) 56–68.

    38. [38]

      R.H. Liang, H.Z. Wu, Z.Z. Hu, et al., Appl. Catal. B: Environ. Energy 352 (2024) 124042. doi: 10.1016/j.apcatb.2024.124042

    39. [39]

      L. Li, J.H. Li, F. Fang, et al., Appl. Catal. B: Environ. Energy 333 (2023) 122776. doi: 10.1016/j.apcatb.2023.122776

    40. [40]

      S.K. Wu, X.F. Wang, H.G. Yu, Chin. J. Struct. Chem. 43 (2024) 100457.

    41. [41]

      J.H. Wu, R.J. Guo, J.W. Wang, et al., Chem. Commun. 60 (2024) 12718–12721. doi: 10.1039/d4cc04436c

    42. [42]

      J.Y. Wang, J.J. Wang, S.J. Zuo, et al., Chin. Chem. Lett. 34 (2023) 108157. doi: 10.1016/j.cclet.2023.108157

    43. [43]

      R.L. Wang, H. Luo, M.D. Sun, et al., J. Catal. 434 (2024) 115521. doi: 10.1016/j.jcat.2024.115521

    44. [44]

      Z.L. Wang, X.X. Duan, M.G. Sendeku, et al., Chem Catal. 3 (2023) 100672.

    45. [45]

      C.F. Chen, C. Wang, Y.N. Zhang, et al., Appl. Catal. B: Environ. Energy 348 (2024) 123854. doi: 10.1016/j.apcatb.2024.123854

    46. [46]

      B.C. Zhu, J.J. Liu, J. Sun, et al., J. Mater. Sci. Technol. 162 (2023) 90–98. doi: 10.1016/j.jmst.2023.03.054

    47. [47]

      S. Siva, G.A. Bodkhe, C.H. Cong, et al., J. Ind. Eng. Chem. 124 (2023) 523–531. doi: 10.1016/j.jiec.2023.05.007

    48. [48]

      S. Zhao, Y.Y. Wan, L. Han, et al., Carbon Lett. 35 (2025) 287–308. doi: 10.1007/s42823-024-00794-2

    49. [49]

      X. Luo, Y.Q. Dong, D.Y. Wang, et al., Rev. Adv. Mater. Sci. 62 (2023) 20230123. doi: 10.1515/rams-2023-0123

    50. [50]

      K. Meng, J.J. Zhang, B. Cheng, et al., Adv. Mater. 36 (2024) 2406460. doi: 10.1002/adma.202406460

    51. [51]

      F. Wei, W.C. Xue, Z.Y. Yu, et al., Chin. Chem. Lett. 35 (2024) 108313. doi: 10.1016/j.cclet.2023.108313

    52. [52]

      Y.L. Niu, C.C. Yue, S.Q. Li, et al., Carbon Lett. 33 (2023) 957–972. doi: 10.1007/s42823-023-00486-3

    53. [53]

      H.X. Liu, W.W. Li, Z. Wu, et al., J. Ind. Eng. Chem. 125 (2023) 95–104. doi: 10.1016/j.jiec.2023.05.016

    54. [54]

      Y.B. Jiang, C. Cao, Y.Y. Tan, et al., J. Mater Sci. Technol. 141 (2023) 32–41. doi: 10.3847/1538-4365/acda89

    55. [55]

      X.T. Xu, S.Q. Dai, S. Xu, et al., Angew. Chem. Int. Ed. 62 (2023) e202309066. doi: 10.1002/anie.202309066

    56. [56]

      Z.J. Gu, Z. Shan, Y.L. Wang, et al., Chin. Chem. Lett. 35 (2024) 108356. doi: 10.1016/j.cclet.2023.108356

    57. [57]

      S. Zhao, S.S. Shen, L. Han, et al., Rare Met. 43 (2024) 4038–4055. doi: 10.1007/s12598-024-02847-x

    58. [58]

      J. Jiang, Z. Xiong, H. Wang, et al., J. Mater. Sci. Technol. 118 (2022) 15–24. doi: 10.1016/j.jmst.2021.12.018

    59. [59]

      J. Zou, G.D. Liao, J.Z. Jiang, et al., Chin. J. Struct. Chem. 41 (2022) 2201025–2201033.

    60. [60]

      Y. Zhou, Y.L. He, M. Gao, et al., Chin. Chem. Lett. 35 (2024) 108690. doi: 10.1016/j.cclet.2023.108690

    61. [61]

      H. Yang, Z.L. Wang, J.F. Zhang, K. Dai, J.X. Low, J. Materiomics 11 (2025) 100996. doi: 10.1016/j.jmat.2024.100996

    62. [62]

      L.F. Chen, A.M.Li H.Y.Zheng, et al., J. Hazard. Mater. 473 (2024) 134634. doi: 10.1016/j.jhazmat.2024.134634

    63. [63]

      C. Liu, M.N. Chen, Y.L. Chen, et al., Chem. Eng. J. 498 (2024) 155646. doi: 10.1016/j.cej.2024.155646

    64. [64]

      Y. Wu, H.P. Xiao, S.X. Jia, et al., J. Alloy. Compd. 1001 (2024) 175171. doi: 10.1016/j.jallcom.2024.175171

    65. [65]

      L.B. Wang, Y.Y. Zhang, X.Y. Wu, et al., J. Alloy. Compd. 1003 (2024) 175671. doi: 10.1016/j.jallcom.2024.175671

    66. [66]

      H.W. Jian, K.M. Deng, T.Y. Wang, et al., Chin. Chem. Lett. 35 (2024) 108651. doi: 10.1016/j.cclet.2023.108651

    67. [67]

      T. Yu, B. Yang, R. Zhang, et al., J. Mater. Sci. Technol. 188 (2024) 11–26. doi: 10.1016/j.jmst.2023.12.004

    68. [68]

      Y. Hu, X.B. Li, W.W. Wang, et al., Chin. J. Struct. Chem. 41 (2022) 2206069–2206078.

    69. [69]

      T. Yu, B. Yang, R. Deng, et al., J. Mater. Chem. A 12 (2024) 13247–13265. doi: 10.1039/d4ta01361a

    70. [70]

      D. Jang, S. Jeon, E.Y. Shin, et al., Carbon Lett. 33 (2023) 803–809. doi: 10.1007/s42823-023-00461-y

    71. [71]

      Z.S. Yang, Q.Q. Zhang, H. Song, et al., Chin. Chem. Lett. 35 (2024) 108418. doi: 10.1016/j.cclet.2023.108418

    72. [72]

      B.R. Li, N. Li, Z.W. Guo, et al., Chem. Eng. J. 500 (2024) 156843. doi: 10.1016/j.cej.2024.156843

    73. [73]

      L. Jian, Y. Dong, H. Zhao, et al., Appl. Catal. B: Environ. Energy 342 (2024) 123340. doi: 10.1016/j.apcatb.2023.123340

    74. [74]

      Y.H. Wang, H. Zhang, D.M. Wei, et al., Chem. Commun. 60 (2024) 10732–10735. doi: 10.1039/d4cc04007d

    75. [75]

      Y.Y. Zhao, J.P. Yuan, L. Zhu, Y. Fang, Chin. Chem. Lett. 35 (2024) 109065. doi: 10.1016/j.cclet.2023.109065

    76. [76]

      S. Yang, A. Verdaguer, L. Arnarson, et al., ACS Catal. 8 (2018) 4064–4081. doi: 10.1021/acscatal.8b00217

    77. [77]

      F.Y. Li, Y. Anjarsari, J.M. Wang, et al., Carbon Lett. 33 (2023) 1321–1331. doi: 10.1007/s42823-022-00380-4

    78. [78]

      W.Q. Chu, H.L. Liu, Q.L. Zhang, et al., J. Ind. Eng. Chem. 125 (2023) 151–162. doi: 10.1016/j.jiec.2023.05.024

    79. [79]

      H. Teymourinia, H.A. Alshamsi, A. Al-nayili, et al., J. Ind. Eng. Chem. 125 (2023) 259–268. doi: 10.1016/j.jiec.2023.05.035

    80. [80]

      X.L. Shao, K. Li, J.P. Li, et al., Chin. J. Catal. 51 (2023) 193–203. doi: 10.1016/S1872-2067(23)64478-X

    81. [81]

      Y. Kofuji, Y. Isobe, Y. Shiraishi, et al., J. Am. Chem. Soc. 138 (2016) 10019–10025. doi: 10.1021/jacs.6b05806

    82. [82]

      H. Hou, X. Zeng, X. Zhang, Angew. Chem. Int. Ed. 59 (2020) 17356–17376. doi: 10.1002/anie.201911609

    83. [83]

      H.C. Wang, Q.X. Zhang, Y.F. Yang, et al., J. Environ. Chem. Eng. 12 (2024) 114979. doi: 10.1016/j.jece.2024.114979

    84. [84]

      X.J. Liu, S. Zhang, M. Wang, J.Q. Wang, Chin. Chem. Lett. 35 (2024) 108723. doi: 10.1016/j.cclet.2023.108723

    85. [85]

      Y.T. Wang, M.X. Wang, X.Y. Zhang, et al., J. Catal. 440 (2024) 115807. doi: 10.1016/j.jcat.2024.115807

    86. [86]

      M. Yang, Y.L. Zou, L. Ding, et al., Carbon Lett. 33 (2023) 1333–1341. doi: 10.1007/s42823-022-00456-1

    87. [87]

      Y.T. Wang, X. Li, T. Mei, et al., Inorg. Chem. Commun. 168 (2024) 112987. doi: 10.1016/j.inoche.2024.112987

    88. [88]

      J.M. Wang, Q. Qin, F.Y. Li, et al., Carbon Lett. 33 (2023) 1381–1394. doi: 10.1007/s42823-022-00401-2

    89. [89]

      Y.H. Wang, W.Y. Yu, C.Y. Wang, et al., eScience 4 (2024) 100228. doi: 10.1016/j.esci.2024.100228

    90. [90]

      Y. Gong, Y. Ding, Q. Tang, et al., Chin. Chem. Lett. 35 (2024) 108475. doi: 10.1016/j.cclet.2023.108475

    91. [91]

      Z.H. Zhang, J.S. Yang, H. Zhao, et al., Surf. Interfaces 55 (2024) 105463. doi: 10.1016/j.surfin.2024.105463

    92. [92]

      E. Gomathi, P. Maharaja, H.S. Rathore, et al., Carbon Lett. 33 (2023) 2011–2025. doi: 10.1007/s42823-023-00551-x

    93. [93]

      R. Xu, Y. Guan, L.X. Cao, et al., J. Water Process Eng. 68 (2024) 106316. doi: 10.1016/j.jwpe.2024.106316

    94. [94]

      J. Li, Y.Q. Feng, F.Y. Fu, et al., Chin. Chem. Lett. 35 (2024) 108736. doi: 10.1016/j.cclet.2023.108736

    95. [95]

      T.L. Yusuf, B.O. Ojo, T. Mushiana, et al., Catal. Sci. Technol. 14 (2024) 6015–6026. doi: 10.1039/d4cy00729h

    96. [96]

      R. Chen, C.K. Chang, Y.L. Hou, et al., Desalin. Water Treat. 319 (2024) 100421. doi: 10.1016/j.dwt.2024.100421

    97. [97]

      C.J. Cai, G.D. Fan, X.F. Cao, et al., J. Hazard. Mater. 470 (2024) 134198. doi: 10.1016/j.jhazmat.2024.134198

    98. [98]

      M.Y. Liu, Y.Y. Wan, Y. Wang, et al., J. Environ. Chem. Eng. 12 (2024) 112328. doi: 10.1016/j.jece.2024.112328

    99. [99]

      S.Y. Li, X.Y. Yang, Q. Wang, et al., Nano Energy 128 (2024) 109866. doi: 10.1016/j.nanoen.2024.109866

    100. [100]

      J.X. Wang, Y.K. Wan, Q.M. He, et al., J. Environ. Chem. Eng. 11 (2023) 111604. doi: 10.1016/j.jece.2023.111604

    101. [101]

      X.B. Li, T. Han, Y.T. Zhou, et al., Appl. Catal. B: Environ. Energy 350 (2024) 123913. doi: 10.1016/j.apcatb.2024.123913

    102. [102]

      S.S. Shen, X.B. Li, Y.T. Zhou, et al., J. Mater. Sci. Technol. 155 (2023) 148–159. doi: 10.1016/j.jmst.2023.03.006

    103. [103]

      X. Ma, H. Cheng, Chem. Eng. J. 429 (2022) 132373. doi: 10.1016/j.cej.2021.132373

    104. [104]

      H. Dong, L. Tong, P. Zhang, et al., J. Mater. Sci. Technol. 179 (2024) 251–261. doi: 10.1016/j.jmst.2023.10.012

    105. [105]

      H. Zhou, W. Ye, J. Jiang, et al., Carbon Lett. 34 (2024) 1569–1591. doi: 10.1007/s42823-024-00748-8

    106. [106]

      M.F. Liang, X.D. Shao, J.Y. Choi, et al., J. Energy Chem. 98 (2024) 714–720. doi: 10.1016/j.jechem.2024.07.007

    107. [107]

      A. Amari, H.S.S. Aljibori, Z. Algarni, et al., J. Ind. Eng. Chem. 140 (2024) 599–616. doi: 10.1016/j.jiec.2024.08.002

    108. [108]

      J. Wang, J. Jiang, F. Li, et al., Green Chem. 25 (2023) 32–58. doi: 10.1039/D2GC03160D

    109. [109]

      L. Lin, Y. Ma, J.J.M. Vequizo, et al., Nat. Commun. 15 (2024) 397. doi: 10.1038/s41467-024-44706-4

    110. [110]

      R.Q. Gao, J.X. Bai, R.C. Shen, et al., J. Mater. Sci. Technol. 137 (2023) 223– 231. doi: 10.1016/j.jmst.2022.09.001

    111. [111]

      Y. Zhang, D. Wen, W. Sun, et al., Chin. J. Struct. Chem. 43 (2024) 100469.

    112. [112]

      R. Shen, N. Li, C. Qin, et al., Adv. Funct. Mater. 33 (2023) 2301463. doi: 10.1002/adfm.202301463

    113. [113]

      F. Li, J. Jiang, J. Wang, et al., Nano Res. 16 (2023) 127–145. doi: 10.1007/s12274-022-4799-z

    114. [114]

      F. Yue, C.T. Wang, W. Duan, et al., Sci. China Chem. 66 (2023) 2109–2120. doi: 10.1007/s11426-023-1636-7

    115. [115]

      T.X. Zeng, Y.N. Hu, Z.Q. Yuan, et al., J. Alloy. Compd. 975 (2024) 172851. doi: 10.1016/j.jallcom.2023.172851

    116. [116]

      X.C. Gao, X. Qu, D.W. Sun, et al., Colloid. Surf. A 656 (2023) 130469. doi: 10.1016/j.colsurfa.2022.130469

    117. [117]

      Y. Zhang, H.S.S. Aljibori, Z. Algarni, et al., Chem. Eng. J. 500 (2024) 156725. doi: 10.1016/j.cej.2024.156725

    118. [118]

      T. Wang, S.S. Qu, J.H. Wang, et al., J. Alloy. Compd. 938 (2023) 168631. doi: 10.1016/j.jallcom.2022.168631

    119. [119]

      C.H. Li, X. Zhang, T. Song, et al., J. Environ. Chem. Eng. 12 (2024) 113396. doi: 10.1016/j.jece.2024.113396

    120. [120]

      T.X. Zhang, T. Zhang, H.P. Fu, et al., J. Solid State. Chem. 331 (2024) 124490. doi: 10.1016/j.jssc.2023.124490

    121. [121]

      X.Y. Li, F. Ye, H. Zhang, et al., J. Environ. Chem. Eng. 11 (2023) 110329. doi: 10.1016/j.jece.2023.110329

    122. [122]

      H.Y. Liu, C.G. Niu, D.W. Huang, et al., Chem. Eng. J. 465 (2023) 143007. doi: 10.1016/j.cej.2023.143007

    123. [123]

      L. Biswal, S.P. Tripathy, S. Dash, et al., Mater. Adv. 5 (2024) 4452–4466. doi: 10.1039/d3ma01176c

    124. [124]

      S.K. Pradhan, R. Bariki, A. Kumar, et al., Surf. Interfaces 52 (2024) 104824. doi: 10.1016/j.surfin.2024.104824

    125. [125]

      H. Chai, J. Nan, W.X. Jin, et al., Chem. Eng. J. 489 (2024) 151293. doi: 10.1016/j.cej.2024.151293

    126. [126]

      J.X. Huang, H.Y. Shi, X.F. Wang, et al., Catal. Sci. Technol. 14 (2024) 2514–2521. doi: 10.1039/d4cy00141a

    127. [127]

      X. Liu, Y.D. Li, K.F. Lin, et al., J. Colloid Interf. Sci. 654 (2024) 1228–1239. doi: 10.1007/s00261-024-04576-2

    128. [128]

      X.B. Li, B.B. Kang, F. Dong, et al., Nano Energy 81 (2021) 105671. doi: 10.1016/j.nanoen.2020.105671

    129. [129]

      X.L. Shao, K. Li, J.P. Li, et al., Chin. J. Catal. 51 (2023) 193–203. doi: 10.1016/S1872-2067(23)64478-X

    130. [130]

      Y. Chang, X.Y. Zhao, Z.Y. Jiang, et al., Chem. Eng. J. 501 (2024) 157717. doi: 10.1016/j.cej.2024.157717

    131. [131]

      M.Z. Shen, X.L. Cai, B.W. Cao, et al., J. Alloy. Compd. 1009 (2024) 176905. doi: 10.1016/j.jallcom.2024.176905

    132. [132]

      H.M. Zhang, H. Bian, F. Wang, et al., Appl. Catal. A: Gen. 686 (2024) 119914. doi: 10.1016/j.apcata.2024.119914

    133. [133]

      Z. Yuan, H. Zhang, J. Jiang, et al., J. Environ. Sci. 156 (2025) 157–172. doi: 10.1016/j.jes.2024.09.007

    134. [134]

      H. Wang, L. Yu, J. Peng, et al., Nano Res. 17 (2024) 8007–8016. doi: 10.1007/s12274-024-6823-y

    135. [135]

      H. Wang, L. Yu, J. Peng, et al., J. Mater. Sci. Technol. 208 (2025) 111–119. doi: 10.1016/j.jmst.2024.05.005

    136. [136]

      S. Wang, D. Zhang, X. Pu, et al., Small 20 (2024) 2311504. doi: 10.1002/smll.202311504

    137. [137]

      J. Jiang, L. Yu, J. Peng, et al., Carbon Lett. 35 (2025) 417–440. doi: 10.1007/s42823-024-00853-8

    138. [138]

      E. Cui, Y. Lu, Z. Li, et al., Chin. Chem. Lett. 36 (2025) 110288. doi: 10.1016/j.cclet.2024.110288

    139. [139]

      Y. Zhang, J. Di, X. Zhu, et al., Appl. Catal. B: Environ. Energy 323 (2023) 122148. doi: 10.1016/j.apcatb.2022.122148

    140. [140]

      X.J. Su, X.Y. Zhang, M.M. Gao, et al., J. Colloid Interf. Sci. 663 (2024) 61–72. doi: 10.1016/j.jcis.2024.02.130

    141. [141]

      J. Wang, Q. Qin, Z. Wang, et al., Acta Phys. Chim. Sin. 40 (2024) 2304044. doi: 10.3866/PKU.WHXB202304044

    142. [142]

      R. Gupta, U. Alam, N. Verma, Chem. Eng. J. 479 (2024) 147644. doi: 10.1016/j.cej.2023.147644

    143. [143]

      L.J. Li, P. Zhang, N. Li, et al., J. Environ. Chem. Eng. 12 (2024) 112142. doi: 10.1016/j.jece.2024.112142

    144. [144]

      S.J. Li, R.Y. Yan, M.J. Cai, et al., J. Mater. Sci. Technol. 164 (2023) 59–67. doi: 10.1016/j.jmst.2023.05.009

    145. [145]

      H. Dong, C. Qu, C. Li, et al., Chin. J. Catal. 70 (2025) 142–206. doi: 10.3390/a18030142

    146. [146]

      S.J. Li, K. Rong, X.Q. Wang, et al., Acta Phys. Chim. Sin. 40 (2024) 2403005. doi: 10.3866/pku.whxb202403005

    147. [147]

      J. Luo, K. Wang, Y.M. Qiu, et al., J. Alloy. Compd. 1008 (2024) 176572. doi: 10.1016/j.jallcom.2024.176572

    148. [148]

      X. Qi, G. Aimaiti, Y. Zou, et al., Appl. Surf. Sci. 688 (2025) 162346. doi: 10.1016/j.apsusc.2025.162346

    149. [149]

      W.L. Fang, L. Wang, X.C. Meng, et al., J. Alloy. Compd. 947 (2023) 169606. doi: 10.1016/j.jallcom.2023.169606

    150. [150]

      Y. Wang, Y.X. He, Y.J. Chi, et al., J. Environ. Chem. Eng. 11 (2023) 109968. doi: 10.1016/j.jece.2023.109968

    151. [151]

      Y.Y. Zhao, C.Y. Yang, S.M. Zhang, et al., Chin. J. Catal. 63 (2024) 258–269. doi: 10.1016/S1872-2067(24)60069-0

    152. [152]

      Z.B. Tong, D.K. Ma, J.A. Wu, et al., Chem. Eng. J. 496 (2024) 154262. doi: 10.1016/j.cej.2024.154262

    153. [153]

      W. Chen, X.B. Han, M. Xu, et al., J. Environ. Chem. Eng. 12 (2024) 114349. doi: 10.1016/j.jece.2024.114349

    154. [154]

      X.Y. Yin, H.Y. Shi, Y. Wang, et al., Acta Phys. Chim. Sin. 40 (2024) 2312007. doi: 10.3866/pku.whxb202312007

    155. [155]

      S. Liu, Q.F. Han, X.H. Liu, J. Alloy. Compd. 1009 (2024) 176965. doi: 10.1016/j.jallcom.2024.176965

    156. [156]

      Y. Zhang, H.X. Xu, Giant 20 (2024) 100335. doi: 10.1016/j.giant.2024.100335

    157. [157]

      S.B. Wu, L.G. Chen, H.Z. He, et al., J. Catal. 438 (2024) 115713. doi: 10.1016/j.jcat.2024.115713

    158. [158]

      L. Hao, L. Kang, H. Huang, et al., Adv. Mater. 31 (2019) e1900546. doi: 10.1002/adma.201900546

    159. [159]

      L. Li, B. Li, L. Feng, et al., Molecules 26 (2021) 3844. doi: 10.3390/molecules26133844

    160. [160]

      T. Sun, C.X. Li, Y.P. Bao, J. Fan, E.Z. Liu, Acta Phys. Chim. Sin. 39 (2023) 2212009. doi: 10.3866/pku.whxb202212009

    161. [161]

      Z. Chen, D. Yao, C. Chu, et al., Chem. Eng. J. 451 (2023) 138489. doi: 10.1016/j.cej.2022.138489

    162. [162]

      Y.T. Li, C.Y. Han, Y. Sui, et al., J. Colloid Interf. Sci. 675 (2024) 560–568. doi: 10.1016/j.jcis.2024.07.037

    163. [163]

      H. Shin, Y. Seo, T. Lee, et al., J. Photoch. Photobio. A 454 (2024) 115725. doi: 10.1016/j.jphotochem.2024.115725

    164. [164]

      C.S. Zhou, L. Tao, J. Gao, et al., J. Environ. Chem. Eng. 12 (2024) 113370. doi: 10.1016/j.jece.2024.113370

    165. [165]

      W.Y. Han, H. Zhang, D.G. Li, et al., Appl. Catal. B: Environ. Energy 350 (2024) 123918. doi: 10.1016/j.apcatb.2024.123918

    166. [166]

      S. Liu, Y. Mao, Q.F. Han, et al., Mat. Sci. Semicon. Proc. 170 (2024) 107978. doi: 10.1016/j.mssp.2023.107978

    167. [167]

      H.G. Liang, H.R. Wang, A.H. Wang, et al., J. Colloid Interf. Sci. 669 (2024) 506–517. doi: 10.1016/j.jcis.2024.04.114

    168. [168]

      X.Q. Li, D.Y. Chen, N.J. Li, et al., J. Colloid Interf. Sci. 633 (2023) 691–702. doi: 10.1016/j.jcis.2022.11.146

    169. [169]

      Y. Guo, X.D. Miao, X.Y. Zhang, et al., Sep. Purif. Technol. 326 (2023) 124807. doi: 10.1016/j.seppur.2023.124807

    170. [170]

      Q.Q. Zhang, H. Miao, J. Wang, et al., Chin. J. Catal. 63 (2024) 176–189. doi: 10.1016/S1872-2067(24)60077-X

    171. [171]

      Z.Y. Liao, J.J. Du, L. Wang, et al., J. Alloy. Compd. 976 (2023) 173322.

    172. [172]

      Y.Y. Chen, L. Zhang, Chem. Eng. J. 495 (2024) 153340. doi: 10.1016/j.cej.2024.153340

    173. [173]

      X.W. Li, K. Yang, F.L. Wang, et al., J. Alloy. Compd. 953 (2023) 170064. doi: 10.1016/j.jallcom.2023.170064

    174. [174]

      H.P. Jiang, X.H. Yu, J.H. Li, et al., Appl. Surf. Sci. 639 (2023) 158281. doi: 10.1016/j.apsusc.2023.158281

    175. [175]

      T. Zhou, X. Liu, L. Zhao, M.T. Qiao, W.Y. Lei, Acta Phys. Chim. Sin. 40 (2024) 2309020. doi: 10.3866/PKU.WHXB202309020

    176. [176]

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

    177. [177]

      T. Liang, Y.T. Li, S. Xu, et al., Chin. J. Catal. 62 (2024) 277–286. doi: 10.1016/S1872-2067(24)60058-6

    178. [178]

      X.T. Ning, Q. Peng, B. Zeng, et al., J. Alloy. Compd. 976 (2024) 173070. doi: 10.1016/j.jallcom.2023.173070

    179. [179]

      G.Q. Chen, Z.X. Zheng, W. Zhong, G.H. Wang, X.H. Wu, Acta Phys. Chim. Sin. 40 (2024) 2406021. doi: 10.3866/pku.whxb202406021

    180. [180]

      W.W. Yang, Y.M. Xu, Q.W. Bu, et al., Appl. Surf. Sci. 643 (2024) 158688. doi: 10.1016/j.apsusc.2023.158688

    181. [181]

      G. Alnaggar, K. Alkanad, S.S.G. Chandrashekar, et al., New J. Chem. 46 (2022) 9629–9640. doi: 10.1039/d2nj00173j

    182. [182]

      J.H. Xia, S.Q. Xing, T. Sun, et al., Renew. Energ. 237 (2024) 121773. doi: 10.1016/j.renene.2024.121773

    183. [183]

      Y. Lu, Y.Y. Wan, J. Liu, et al., Sep. Purif. Technol. 358 (2025) 130126. doi: 10.1016/j.seppur.2024.130126

    184. [184]

      J.H. Xia, T. Gao, H.X. Ma, et al., Ceram. Int. 50 (2024) 48814–48825. doi: 10.1016/j.ceramint.2024.09.235

    185. [185]

      J.J. Liu, Y.D. Fu, G.L. Chu, et al., Process Saf. Environ. 191 (2024) 883–896. doi: 10.1016/j.psep.2024.09.040

    186. [186]

      A.H. Bhosale, S. Narra, S.S. Bhosale, E.W. Diau, J. Phys. Chem. Lett. 13 (2022) 7987–7993. doi: 10.1021/acs.jpclett.2c02153

    187. [187]

      L.T. Wu, D.Y. Yang, Y.L. Dong, et al., Chem. Eng. J. 500 (2024) 156777. doi: 10.1016/j.cej.2024.156777

    188. [188]

      M.L. Li, H. Cui, Y. Zhao, et al., Catalysts 14 (2024) 374. doi: 10.3390/catal14060374

    189. [189]

      A. Xu, Y.K. Zhang, H.G. Fan, et al., ACS Appl. Nano Mater. 7 (2024) 3488–3498. doi: 10.1021/acsanm.4c00147

    190. [190]

      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

    191. [191]

      M. Madkour, Y. Abdelmonem, U.Y. Qazi, et al., RSC Adv. 11 (2021) 29433–29440. doi: 10.1039/d1ra04649g

    192. [192]

      F.H. Li, X.H. Li, S.L. Tong, et al., Nano Energy 117 (2023) 108849. doi: 10.1016/j.nanoen.2023.108849

    193. [193]

      Y.Q. Zhi, J.L. Tian, J.H. Sun, et al., J. Alloy. Compd. 965 (2023) 171377. doi: 10.1016/j.jallcom.2023.171377

    194. [194]

      W.T. Huo, M.L. Wang, H. Wei, et al., Appl. Surf. Sci. 639 (2023) 158199. doi: 10.1016/j.apsusc.2023.158199

    195. [195]

      P.Y. Hao, Y.L. Cao, X.E. Ning, et al., J. Colloid Interf. Sci. 639 (2023) 343– 354. doi: 10.1016/j.jcis.2023.02.075

    196. [196]

      Y. Zhang, J. Di, X. Zhu, et al., Appl. Catal. B: Environ. Energy 323 (2023) 122148. doi: 10.1016/j.apcatb.2022.122148

    197. [197]

      A. Kumar, S. Rana, T.T. Wang, et al., Mat. Sci. Semicon. Proc. 168 (2023) 107869. doi: 10.1016/j.mssp.2023.107869

    198. [198]

      Z.C. Zhao, K. Wang, L. Chang, et al., Sol. Energy 264 (2023) 112042. doi: 10.1016/j.solener.2023.112042

    199. [199]

      C. Zuo, X.S. Tai, Q. Su, et al., Opt. Mater. 137 (2023) 113560. doi: 10.1016/j.optmat.2023.113560

    200. [200]

      S.D. Yuan, J.F. Wang, C.R. Zhao, et al., Sep. Purif. Technol. 325 (2023) 124665. doi: 10.1016/j.seppur.2023.124665

    201. [201]

      Y. Yang, J.J. Liu, M.L. Gu, et al., Appl. Catal. B: Environ. Energy 333 (2023) 122780. doi: 10.1016/j.apcatb.2023.122780

    202. [202]

      Y. Xia, B.C. Zhu, X. Qin, et al., Chem. Eng. J. 467 (2023) 143528. doi: 10.1016/j.cej.2023.143528

    203. [203]

      Y. Xia, K.Y. Zhang, H. Yang, et al., Acta Phys. Chim. Sin. 40 (2024) 2407012. doi: 10.3866/pku.whxb202407012

    204. [204]

      C.H. Xia, L. Yuan, H. Song, et al., Small 19 (2023) 2300292. doi: 10.1002/smll.202300292

    205. [205]

      K.Y. Zhang, Y.F. Li, S.D. Yuan, et al., Acta Phys. Chim. Sin. 39 (2023) 2212010. doi: 10.3866/pku.whxb202212010

    206. [206]

      L. Wang, J. Sun, B. Cheng, et al., J. Phys. Chem. Lett. 14 (2023) 4803–4814. doi: 10.1021/acs.jpclett.3c00811

    207. [207]

      L.W. Wei, S.H. Liu, H.P. Wang, et al., ACS Appl. Mater. Interfaces 15 (2023) 25473–25483. doi: 10.1021/acsami.3c02383

    208. [208]

      R. Bariki, S.K. Pradhan, S. Panda, et al., Langmuir 22 (2023) 7707–7722. doi: 10.1021/acs.langmuir.3c00519

    209. [209]

      D.L. Liang, Z.L. Lin, Y.L. Wu, et al., Environ. Sci.: Nano 10 (2023) 1053–1064. doi: 10.1039/d3en00056g

    210. [210]

      X. Han, T.D. Zhang, Y. Cui, et al., Molecules 28 (2023) 2422. doi: 10.3390/molecules28062422

    211. [211]

      J.H. Baek, T.M. Gill, H. Abroshan, et al., ACS Energy Lett. 4 (2019) 720–728. doi: 10.1021/acsenergylett.9b00277

    212. [212]

      H. Yu, X.N. Zhang, Q. Chen, et al., Chem. Res. Chin. Univ. 41 (2025) 734–740. doi: 10.1007/s40242-024-4213-3

    213. [213]

      Y. Xue, Y. Wang, Z. Pan, et al., Angew. Chem. Int. Ed. 60 (2021) 10469–10480. doi: 10.1002/anie.202011215

    214. [214]

      C.D. Han, Y.C. Zhang, Q. Zhang, et al., Rare Met. 43 (2023) 500–510.

    215. [215]

      Y.S. Zhang, H.M. Duan, N. Wang, et al., Chem. Eng. J. 477 (2023) 146903. doi: 10.1016/j.cej.2023.146903

    216. [216]

      L.J. Yang, A. Zhang, L. Zhang, et al., ACS Appl. Mater. Interfaces 15 (2023) 18951–18961. doi: 10.1021/acsami.3c00578

    217. [217]

      A. Hayat, M. Sohail, S.B. Moussa, et al., Adv. Colloid Interfaces 319 (2023) 102969. doi: 10.1016/j.cis.2023.102969

    218. [218]

      K.P. Liu, N. Wang, J.H. Li, et al., J. Ind. Eng. Chem. 128 (2023) 586–596. doi: 10.1016/j.jiec.2023.08.025

    219. [219]

      Y.L. Zang, Z.H. Kang, Y.Z. Gong, et al., Appl. Surf. Sci. 657 (2024) 159594. doi: 10.1016/j.apsusc.2024.159594

    220. [220]

      R. Kaushik, K. Sharma, P.F. Siril, et al., Chem. Eng. J. 479 (2024) 147575. doi: 10.1016/j.cej.2023.147575

    221. [221]

      Y. Wu, C. Cheng, K.Z. Qi, et al., Acta Phys. Chim. Sin. 40 (2024) 2406027. doi: 10.3866/pku.whxb202406027

    222. [222]

      Y. Zhang, J.Y. Qiu, B.C. Zhu, et al., Chem. Eng. J. 444 (2022) 136584. doi: 10.1016/j.cej.2022.136584

    223. [223]

      W. Gan, J. Guo, X.C. Fu, et al., Sep. Purif. Technol. 317 (2023) 123791. doi: 10.1016/j.seppur.2023.123791

    224. [224]

      Z.C. Jiang, Q. Long, B. Cheng, et al., J. Mater. Sci. Technol. 162 (2023) 1–10. doi: 10.1016/j.jmst.2023.03.045

    225. [225]

      X.B. Li, J. Xiong, X.M. Gao, et al., J. Hazard. Mater. 387 (2020) 121690. doi: 10.1016/j.jhazmat.2019.121690

    226. [226]

      K. Li, C. Liu, J. Li, et al., Acta Phys. Chim. Sin. 40 (2024) 2403009. doi: 10.3866/pku.whxb202403009

    227. [227]

      X.B. Li, J.Y. Liu, J.T. Huang, et al., Acta Phys. Chim. Sin. 37 (2021) 2010030.

    228. [228]

      Y.A. Liu, J.F. You, Q.X. Shi, et al., Appl. Catal. A: Gen. 689 (2025) 120005. doi: 10.1016/j.apcata.2024.120005

    229. [229]

      L.F. Ning, X. Chen, Z.P. Wang, et al., Appl. Catal. B: Environ. Energy 324 (2023) 122282. doi: 10.1016/j.apcatb.2022.122282

    230. [230]

      J.R. Han, Z. Wang, X.Y. Yang, et al., Appl. Catal. A: Gen. 687 (2024) 119967. doi: 10.1016/j.apcata.2024.119967

    231. [231]

      L.H. Meng, C. Zhao, X. Zhang, et al., Nano Energy 128 (2024) 109795. doi: 10.1016/j.nanoen.2024.109795

    232. [232]

      Y. Wu, Y. Yang, M.L. Gu, et al., Chin. J. Catal. 53 (2023) 123–133. doi: 10.1016/S1872-2067(23)64514-0

    233. [233]

      S. Zhou, D. Wen, W. Zhong, et al., J. Mater. Sci. Technol. 199 (2024) 53–65. doi: 10.1016/j.jmst.2024.02.048

    234. [234]

      Y.Y. Chen, L. Zhang, Appl. Catal. B: Environ. Energy 347 (2024) 123768. doi: 10.1016/j.apcatb.2024.123768

    235. [235]

      D.L. Dai, J.H. Qiu, G.L. Xia, et al., Small 20 (2024) 2403268.

    236. [236]

      B.W. Deng, P. Chen, P. Xie, et al., Chem. Eng. Sci. 267 (2023) 118368. doi: 10.1016/j.ces.2022.118368

    237. [237]

      X.Q. Zhang, J.M. Liu, R. Li, et al., J. Colloid Interf. Sci. 645 (2023) 956–963. doi: 10.1016/j.jcis.2023.05.011

    238. [238]

      A. Balakrishnan, M. Chinthala, A. Kumar, et al., Chem. Eng. J. 496 (2024) 153899. doi: 10.1016/j.cej.2024.153899

    239. [239]

      Z.Q. Zan, X.B. Li, X. Gao, et al., Acta Phys. Chim. Sin. 39 (2023) 2209016.

    240. [240]

      L. Li, B. Li, L. Feng, et al., Molecules 26 (2021) 3844. doi: 10.3390/molecules26133844

    241. [241]

      H. Guo, L. Zhou, K. Huang, et al., Adv. Funct. Mater. 34 (2024) 2402650. doi: 10.1002/adfm.202402650

    242. [242]

      X. Zhang, Y. Zeng, W. Shi, et al., Chem. Eng. J. 429 (2022) 131312. doi: 10.1016/j.cej.2021.131312

    243. [243]

      Y. Li, Y. Guo, G. Fan, et al., Angew. Chem. Int. Ed. 63 (2024) e202317572. doi: 10.1002/anie.202317572

    244. [244]

      R. Ji, Y.M. Dong, H. Zhao, et al., J. Colloid Interf. Sci. 648 (2023) 623–632. doi: 10.1016/j.jcis.2023.05.180

    245. [245]

      X. Feng, L. Zhang, J. Mater. Sci. Technol. 156 (2023) 54–63. doi: 10.1016/j.jmst.2022.09.040

    246. [246]

      Y. Sun, L. Zhang, Appl. Catal. B: Environ. Energy 343 (2024) 123549. doi: 10.1016/j.apcatb.2023.123549

    247. [247]

      S. Jing, J. Zhao, A. Wang, et al., Chem. Eng. J. 479 (2024) 147150. doi: 10.1016/j.cej.2023.147150

    248. [248]

      H. Du, R.B. Huo, A.X. Liu, et al., Appl. Catal. B: Environ. Energy 357 (2024) 124283. doi: 10.1016/j.apcatb.2024.124283

    249. [249]

      K.Q. Sun, H. Yuan, Y.J. Yan, et al., J. Water Process Eng. 58 (2024) 104803. doi: 10.1016/j.jwpe.2024.104803

    250. [250]

      Y. Xu, J. Liao, L. Zhang, et al., J. Colloid Interf. Sci. 647 (2023) 446–455. doi: 10.1016/j.jcis.2023.05.140

    251. [251]

      D. Prusty, S. Mansingh, K.M. Parida, Catal. Sci. Technol. 13 (2023) 1311–1324. doi: 10.1039/d2cy02107b

    252. [252]

      W. Yu, Z. Zhu, C. Hu, et al., J. Mater. Chem. A 11 (2023) 6384–6393. doi: 10.1039/d2ta10074f

    253. [253]

      Y. Yang, Q. Wang, X. Zhang, et al., J. Mater. Chem. A 11 (2023) 1991–2001. doi: 10.1039/d2ta08145h

    254. [254]

      L. Yang, J. Wang, T. Ma, et al., J. Colloid Interf. Sci. 611 (2022) 760–770. doi: 10.1016/j.jcis.2021.11.100

    255. [255]

      H.Z. Zhang, J.J. Liu, Y. Zhang, et al., J. Mater. Sci. Technol. 166 (2023) 241–249. doi: 10.1016/j.jmst.2023.05.030

    256. [256]

      X. Zhang, J. Yu, W. Macyk, et al., Adv. Sustain. Syst. 7 (2022) 2200113.

    257. [257]

      C. Lai, M. Xu, F. Xu, et al., Chem. Eng. J. 452 (2023) 139070. doi: 10.1016/j.cej.2022.139070

    258. [258]

      S.J. Li, C.C. Wang, K.X. Dong, et al., Chin. J. Catal. 51 (2023) 101–112.

    259. [259]

      K. Li, J.L. Mei, J.P. Li, et al., Sci. China Mater. 67 (2024) 484–492. doi: 10.1007/s40843-023-2717-0

    260. [260]

      W.W. Wang, X.B. Li, F. Deng, et al., Chin. Chem. Lett. 33 (2022) 5200–5207. doi: 10.1016/j.cclet.2022.01.058

    261. [261]

      G. Ding, Z. Wang, J. Zhang, et al., EcoEnergy 2 (2024) 22–44. doi: 10.1002/ece2.25

    262. [262]

      Y. Dong, B. Wang, D. Xie, et al., EcoEnergy 2 (2024) 489–502. doi: 10.1002/ece2.54

    263. [263]

      X.B. Li, T. Han, Y.T. Zhou, et al., Sci. China Technol. Sci. 67 (2024) 1238–1252. doi: 10.1007/s11431-023-2604-x

    264. [264]

      L.F. Xiao, W.L. Ren, S.S. Shen, et al., Acta Phys. Chim. Sin. 40 (2024) 2308036. doi: 10.3866/PKU.WHXB202308036

    265. [265]

      X.B. Li, H.S. Zhang, J.M. Luo, et al., Electrochim. Acta 258 (2017) 998–1007. doi: 10.1016/j.electacta.2017.11.151

    266. [266]

      Y. Wang, Q. Liu, N.H. Wong, et al., Ceram. Int. 48 (2022) 2459–2469. doi: 10.1016/j.ceramint.2021.10.027

    267. [267]

      S.J. Li, X.B. Chen, Y.P. Yuan, Acta Phys. Chim. Sin. 39 (2023) 2303032. doi: 10.3866/pku.whxb202303032

    268. [268]

      K. Wang, Q.P. Wang, K.J. Zhang, et al., J. Mater. Sci. Technol. 124 (2022) 202–208. doi: 10.1016/j.jmst.2021.10.059

    269. [269]

      Z. Chen, D.C. Yao, C.C. Chu, et al., Chem. Eng. J. 451 (2023) 138489. doi: 10.1016/j.cej.2022.138489

    270. [270]

      X. Fang, X. Huang, Q.Y. Hu, et al., Chem. Commun. 60 (2024) 5354. doi: 10.1039/d4cc01577k

    271. [271]

      X.T. Xu, C.F. Shao, J.F. Zhang, Z.L. Wang, K. Dai, Acta Phys. Chim. Sin. 40 (2024) 2309031. doi: 10.3866/PKU.WHXB202309031

    272. [272]

      J.Z. Cheng, B. Cheng, J.S. Xu, J.G. Yu, S.W. Cao, eScience 5 (2025) 100354. doi: 10.1016/j.esci.2024.100354

    273. [273]

      H.W. He, Z.L. Wang, J.F. Zhang, et al., Adv. Funct. Mater. 34 (2024) 2315426. doi: 10.1002/adfm.202315426

    274. [274]

      B.C. Zhu, C.J. Jiang, J.S. Xu, et al., Mater. Today 82 (2025) 251–273. doi: 10.1016/j.mattod.2024.11.012

    275. [275]

      J.H. Hua, Z.L. Wang, J.F. Zhang, et al., J. Mater. Sci. Technol. 156 (2023) 64–71. doi: 10.1016/j.jmst.2023.03.003

    276. [276]

      L.X. Chen, S.Y. Qiu, Y.J. Cai, et al., Sep. Purif. Technol. 358 (2025) 130209. doi: 10.1016/j.seppur.2024.130209

    277. [277]

      J.Q. Ye, Y.Y. Wan, Y.J. Li, et al., Appl. Surf. Sci. 684 (2025) 161862. doi: 10.1016/j.apsusc.2024.161862

    278. [278]

      J. Chen, L. Hu, Q.H. Chen, et al., Surf. Interfaces 46 (2024) 104184. doi: 10.1016/j.surfin.2024.104184

    279. [279]

      C. Chen, F.T. Liao, X.C. Zhang, et al., RSC Adv. 14 (2024) 12407–12415. doi: 10.1039/d4ra01746c

    280. [280]

      J.T. Wang, M.C. Gao, M.Y. Xu, et al., Colloid. Surf. A 702 (2024) 135169. doi: 10.1016/j.colsurfa.2024.135169

    281. [281]

      Y.Q. Wang, Z.Y. Fan, Y.Y. Wan, et al., Appl. Catal. B: Environ. Energy 366 (2025) 125017. doi: 10.1016/j.apcatb.2024.125017

    282. [282]

      Y.Y. Fu, Y.X. Xie, Y. Zhao, et al., Eur. Polym. J. 210 (2024) 112945. doi: 10.1016/j.eurpolymj.2024.112945

    283. [283]

      Y.P. Cui, H.D. Yang, X. Xiang, et al., Chem. Eng. J. 489 (2024) 151445. doi: 10.1016/j.cej.2024.151445

    284. [284]

      R.Z. Li, J.D. Luan, Y. Zhang, et al., Renew. Sust. Energ. Rev. 206 (2024) 114863. doi: 10.1016/j.rser.2024.114863

    285. [285]

      J.Q. Ye, S.Y. Xu, Y.Y. Wan, J. Colloid Interf. Sci. 685 (2025) 304–320. doi: 10.1016/j.jcis.2025.01.105

    286. [286]

      M.V.D. Prado, V. Lima, L. Oliveira, et al., Appl. Clay Sci. 262 (2024) 107622. doi: 10.1016/j.clay.2024.107622

    287. [287]

      K.D. Jayeola, D.S. Sipuka, T.I. Sebokolodi, et al., Electrochim. Acta 507 (2024) 145160. doi: 10.1016/j.electacta.2024.145160

    288. [288]

      S.S. Zhi, X.X. Zou, C. Yang, D.P. Wu, H. Guo, Appl. Catal. B: Environ. Energy 371 (2025) 125215. doi: 10.1016/j.apcatb.2025.125215

    289. [289]

      L. Zhang, H.T. Wang, J. Peng, et al., Sep. Purif. Technol. 364 (2025) 132284. doi: 10.1016/j.seppur.2025.132284

    290. [290]

      B.B. Li, C.Y. Ning, Y.L. Pan, et al., J. Environ. Chem. Eng. 13 (2025) 115301. doi: 10.1016/j.jece.2024.115301

    291. [291]

      C. Yang, Y. Xiang, W. Wang, et al., Appl. Catal. B: Environ. Energy 365 (2025) 124856. doi: 10.1016/j.apcatb.2024.124856

    292. [292]

      X.D. Liu, Y.F. Li, H.L. Wang, et al., Carbon Lett. 34 (2024) 1995–2011. doi: 10.1007/s42823-024-00741-1

    293. [293]

      T. Zhang, Z.J. Song, Z.W. Sun, et al., J. Colloid Interf. Sci. 686 (2025) 1114–1124. doi: 10.1016/j.jcis.2025.02.014

    294. [294]

      Y.Y. Gong, Y.L. Cao, Y.Y. Feng, et al., Mater. Lett. 379 (2025) 137617. doi: 10.1016/j.matlet.2024.137617

    295. [295]

      M.A. Khlifi, W.H. Hassan, A.A.H. Kadhum, et al., J. Water Process Eng. 71 (2025) 107111. doi: 10.1016/j.jwpe.2025.107111

    296. [296]

      N. Sun, J. Chen, Z.L. Zhang, et al., Fuel 386 (2025) 134185. doi: 10.1016/j.fuel.2024.134185

    297. [297]

      Y.N. Wang, Y.X. Guo, J.S. Peng, et al., J. Environ. Chem. Eng. 13 (2025) 115296. doi: 10.1016/j.jece.2024.115296

    298. [298]

      J.N. Gao, Z.P. Guo, Y.H. Li, B. Yang, A. Wei, Chem. Eng. J. 500 (2024) 156852. doi: 10.1016/j.cej.2024.156852

    299. [299]

      L. Ding, M.J. Lei, T. Wang, J. Wang, Z.L. Jin, Carbon Lett. 34 (2024) 2099–2112. doi: 10.1007/s42823-024-00743-z

    300. [300]

      J.N. Qu, Y.H. Zhang, X.L. Liang, Sep. Purif. Technol. 359 (2025) 130813. doi: 10.1016/j.seppur.2024.130813

    301. [301]

      T. Fazliev, D. Polskikh, D. Selishchev, Appl. Surf. Sci. 686 (2025) 162124. doi: 10.1016/j.apsusc.2024.162124

  • Figure 1  Summary of this review content.

    Figure 2  Schematic diagram of relevant mechanisms of the photocatalytic H2O2 production.

    Figure 3  (a) Preparation route of CdS/RF composite. Copied with permission [46]. Copyright 2023, Elsevier. (b) The synthesis process for iCOF/BO composite. Copied with permission [203]. Copyright 2024, Editorial Office of Acta Physico-Chimica Sinica. (c) Synthetic procedure of TiO2@BTTA composites. Copied with permission [201]. Copyright 2023, Elsevier.

    Figure 4  (a) Synthesis route of ZnO/PBD. Copied with permission [221]. Copyright 2024, Editorial Office of Acta Physico-Chimica Sinica. (b) Formation illustration of WS2/S-g-C3N4 Z-scheme heterostructures. Copied with permission [119]. Copyright 2024, Elsevier. (c) Schematic representation of preparing ZIS-Z/OCN. Copied with permission [122]. Copyright 2023, Elsevier.

    Figure 5  (a) Schematic illustration of the formation of the ZT heterojunction. Copied with permission [222]. Copyright 2022, Elsevier. (b) Schematic representation of the synthesis process for 3DOM SCN/T. Copied with permission [224]. Copyright 2023, Elsevier. (c) Schematic depiction of the preparation process for Sol-BP/BiOBr composites. Copied with permission [225]. Copyright 2020, Elsevier.

    Figure 6  (a) The fabrication process of PDI-Ala/S-C3N4. Copied with permission [227]. Copyright 2021, Editorial Office of Acta Physico-Chimica Sinica. (b) Synthetic pathways of CoPc/K/Na/PCN. Copied with permission [228]. Copyright 2024, Elsevier.

    Figure 7  (a) Schematic representation of the fundamental synthetic protocol for ZCx. Copied with permission [231]. Copyright 2024, Elsevier. (b) Schematic illustration of the synthesis process for ZnO/ZnIn2S4. Copied with permission [232]. Copyright 2023, Elsevier. (c) Schematics for the synthesis of ZnO/CuInS2 photocatalyst. Copied with permission [50]. Copyright 2024, Wiley-VCH.

    Figure 8  (a) The synthetic process of the TCNx sample. Copied with permission [233]. Copyright 2024, Elsevier. (b) The illustration to prepare In2S3/MnIn2S4 catalyst. Copied with permission [234]. Copyright 2024, Elsevier. (c) Schematic showing the synthesis process of CN/HMoP synthesis. Copied with permission [127]. Copyright 2024, Elsevier. (d) Creative synthetic method of ZnIn2S4/Bi2S3 hybrids in this work. Copied with permission [235]. Copyright 2024, Wiley-VCH.

    Figure 9  (a) Scheme of iML method. Copied with permission [236]. Copyright 2022, Elsevier. (b) Selection methodology for SACs exhibiting elevated selectivity and catalytic activity. Copied with permission [237]. Copyright 2023, Elsevier.

    Figure 10  ESR detection for DMPO-O2 (a), DMPO-OH (b) and TEMPO-h+ (c) under dark and visible-light irradiation, schematic illustrations of WO2.72 and 0.5S-pCN before and after contact, as well as the formation of IEF and band bending (d). Copied with permission [128]. Copyright 2023, Elsevier. The photocatalytic H2O2 production yield of 40ZIS-Z/OCN under various conditions (e), ESR spectra of DMPO-OH (f), DMPO-O2 (g) and TEMP-1O2 (h) for 40ZIS-Z/OCN under visible light irradiation. Copied with permission [122]. Copyright 2023, Elsevier. Transient photocurrent response (i), EIS analysis (j) and K-L curves (k) of ZnO, ZnIn2S4, and ZZS-20. Copied with permission [232]. Copyright 2023, Elsevier.

    Figure 11  In-situ FTIR spectra of a series of OPACN composites during the photocatalytic H2O2 production (a) and an enlarged view of the FTIR peak (b, c). Copied with permission [238]. Copyright 2024, Elsevier. (d-f) In-situ XPS spectra of C, N and Bi in CNQDs/BiOBr-1.50%. Copied with permission [239]. Copyright 2023, Editorial Office of Acta Physico-Chimica Sinica. (g) In-situ ESR spectra of 0.1% Au/TiO2 composites and pure TiO2. Copied with permission [240], Copyright 2021, MDPI. (h) In-situ DRIFTS of CDs10MCN collected in the H2O2 photocatalytic processes in the O2 atmosphere. Copied with permission [241]. Copyright 2024, Wiley-VCH. In-situ KPFM potential diagram in dark (i), under light irradiation (j), and surface potential (k) of PCN(5)/MnS-D. Copied with permission [150]. Copyright 2023, Elsevier.

    Figure 12  Optimal configurations of H2O2 adsorption on TiO2 (101) (a) and 2H-MoSe2 (103) planes (b). Copied with permission [242]. Copyright 2022, Elsevier. Optimized unit cells of states for CDs10MCN (c) and free energy diagram of CDs10MCN for photocatalytic H2O2 production (d). Copied with permission [241]. Copyright 2024, Wiley-VCH. (e-k) The theoretical calculation results regarding the Zn-N4 and Zn-N3O models. Copied with permission [243]. Copyright 2024, Wiley-VCH.

    Figure 13  Photoluminescence spectra (a), EIS Nyquist plot (b), Mott-schottky plot (c) and the yield of H2O2 (d) for the catalysts. (e) The results of repeatability experiments on the production of H2O2 by the OPACN hydrogel. (f) The photocatalytic mechanism illustration of OPACN hydrogel. Copied with permission [238]. Copyright 2024, Elsevier.

    Figure 14  (a) H2O2 production of the samples. (b) The comparison of TOC removal rate. (c) The TOC removal rate of various organic pollutants over CoFeO/PDIsm. (d) The rate constants of the reaction for the degradation of CIP under diverse systems. (e) The concentration of Fe2+ under diverse systems. (f) The concentration of radical OH under diverse systems (f). (g) OH transformation rate under diverse systems. (h) The postulated mechanism for the degradation reaction induced by CoFeO/PDIsm. Copied with permission [244]. Copyright 2023, Elsevier.

    Figure 15  The selective photocatalytic oxidation of benzaldehyde over various samples (a), the HPLC chromatogram depicting benzaldehyde and the resultant benzoin dimethyl acetal generated within YS-CuCo2S4@Cu2O-NR (b), time-dominated selective oxidation of benzaldehyde (c), the recyclability of YS-CuCo2S4@Cu2O-NR during the oxidation of benzaldehyde accompanied by the production of H2O2 (d), EPR signals recorded for YS-CuCo2S4@ Cu2O-NR under light illumination (e), DMPO-C/O2− in benzaldehyde with YS-CuCo2S4@Cu2O-NR (f). Copied with permission [245]. Copyright 2024, Elsevier.

    Figure 16  LSV curves of different electrodes (a), the influence of different applied potentials on the production of H2O2 (b), H2O2 yields and FEs of different electrodes (c), stability experiment for the H2O2 production via PEC (d), H2O2 yield and FEs of the Janus electrode (e), the influence of different conditions on As(Ⅲ) oxidation and ROX degradation (f), both the conversion efficiency of As(Ⅲ) and the degradation efficiency of ROX over different systems (g), schematic illustration of the mechanism underlying the photoelectrochemical process (h). Copied with permission [246]. Copyright 2024, Elsevier.

    Figure 17  Photocatalytic H2O2 production by the samples (a), degradation performance (b) and the corresponding degradation kinetics (c) of TC by the samples, PL (d), TRPL (e) and EIS (f) spectra, along with photocurrent response (g) of different samples, mechanism diagram of the coupling of H2O2 production and TC degradation by photocatalysis of MCN@CdS S-scheme heterojunction composite (h). Copied with permission [247]. Copyright 2024, Elsevier.

    Figure 18  TC degradation efficiency of diverse samples (a), the cell density of E. coli (b) and S. aureus (c) for diverse samples, the antibacterial performances of diverse samples for E. coli (d) and S. aureus (e). Copied with permission [248]. Copyright 2024, Elsevier. Concentration of H2O2 production curves for different samples (f), the TC degradation performance upon diverse conditions (g), possible mechanism for TC degradation of BFO/ZIS (h). Copied with permission [249]. Copyright 2024, Elsevier.

    Figure 19  H2O2 production at the anode (a), LSV plots of In2S3/MnIn2S4/PVDF/NF||In2S3/MnIn2S4/CP (b), different potentials for the production of H2O2 and HClO (c), recyclability and stability of ORR||ClOR system (d), degradation efficiency of MB in different systems (e), mechanism of OH activation (f), mechanism diagram of organic pollutant degradation by H2O2/HClO production in In2S3/MnIn2S4 photoelectrocatalytic system (g). Copied with permission [234]. Copyright 2024, Elsevier.

    Figure 20  A schematic diagram of the research directions and development prospects of photocatalytic and photoelectrocatalytic H2O2 production.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  79
  • HTML全文浏览量:  0
文章相关
  • 发布日期:  2025-10-15
  • 收稿日期:  2025-02-08
  • 接受日期:  2025-06-06
  • 修回日期:  2025-05-29
  • 网络出版日期:  2025-06-07
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