Surface engineering strategies for particulate photocatalysts toward photocatalytic H2O2 production

Yu Zhang a, Zhian Chenb, Menghe Luoa, Sheng Ye*b and Shanshan Chen*a
aSchool of Materials Science and Engineering, Nankai University, Tianjin 300350, China. E-mail: sschen@nankai.edu.cn
bAgricultural Photocatalysis Laboratory, School of Materials and Chemistry, Anhui Agricultural University, Hefei 230036, China. E-mail: sye503@ahau.edu.cn

Received 14th April 2025 , Accepted 24th June 2025

First published on 26th June 2025


Abstract

The photocatalytic production of hydrogen peroxide (H2O2) using particulate photocatalysts is a safe, sustainable and green process that requires only oxygen and water as feedstocks and solar energy as a power source. Surface engineering on particulate photocatalysts can significantly improve the performance of photocatalytic H2O2 production by optimizing light absorption, surface charge separation, and reaction pathways. To provide a comprehensive and systematic illustration of this topic, various surface engineering strategies are classified and elaborated in this review, which are mainly included in the following two aspects. The first one involves surface modification relating to crystal facets, surface vacancies, and surface functional groups. The second one focuses on surface composite strategies, including combination with metals, semiconductors, carbon nanomaterials, polyoxometalates or their derivatives, and organic compounds. Finally, the challenges and prospects in the surface engineering strategies for particulate photocatalysts for promoting photocatalytic H2O2 production are analyzed and discussed.



Wider impact

Photocatalytic H2O2 production has attracted extensive attention in recent years because it uses only H2O and O2 as source materials and can be operated under ambient conditions. In a typical photocatalyst system, the surface of a particulate semiconductor provides an important platform for the adsorption and activation of reactant molecules, the separation and transfer of surface charge carriers, the promotion of H2O2 formation and the prohibition of H2O2 decomposition. Therefore, innovative surface engineering strategies are summarized and illustrated in this review, aiming to provide a deep understanding of the structure-performance relationship and construct efficient photocatalytic H2O2 production systems. Specifically, surface engineering strategies, including surface modifications and surface composite strategies, are classified and elaborated. The challenges and prospects in the surface engineering strategies for particulate photocatalysts are further analyzed and summarized. This review is anticipated to offer valuable guidance for the design and fabrication of advanced particulate photocatalysts for efficient H2O2 synthesis.

1. Introduction

H2O2 has attracted considerable scientific interest since its synthesis was initiated in the eighteenth century and has been recognized as one of the 100 most important chemicals globally.1–3 H2O2 has strong oxidation ability, and its decomposition generates water (H2O) and oxygen (O2) as the only by-products without any pollution; this makes it an ideal green oxidizer for environmental remediation and the chemical industry.4 Furthermore, owing to its high energy density, H2O2 has garnered significant attention in the field of energy conversion.5,6 Compared with molecular hydrogen (H2), H2O2 has demonstrated superiority in enabling liquid-phase storage and transport under ambient conditions. These characteristics render H2O2 as a promising alternative energy carrier, exhibiting distinct advantages over H2 in energy storage and transport systems.7,8

Currently, the most widely used method for industrial H2O2 production is the anthraquinone process. As shown in Fig. 1, this process comprises four sequential steps: (1) anthraquinones (e.g., 2-ethylanthraquinone) undergo hydrogenation with H2, forming hydrogenated anthraquinones. (2) Hydrogenated derivatives are oxidized in the presence of O2, yielding H2O2. (3) The generated H2O2 is extracted via H2O to produce an aqueous solution. (4) The regenerated anthraquinones are recycled into subsequent hydrogenation and oxidation cycles, creating a closed-loop system.9 Although the process is widely used, the high energy consumption of the hydrogenation step and the associated organic waste discharge problem result in low overall energy efficiency and high environmental risks. These inherent limitations highlight the critical demand for eco-friendly and sustainable alternatives to H2O2 synthesis.10


image file: d5mh00697j-f1.tif
Fig. 1 Schematic representation of the anthraquinone oxidation process. AQ, AHQ, THAHQ, and THAQ represent 2-alkylanthraquinone, 2-alkylanthrahydroquinone, 5,6,7,8-tetrahydroanthrahydroquinone, and 5,6,7,8-tetrahydroanthraquinone, respectively.

Amidst escalating global energy demands and resource shortages, the development of renewable energy sources, particularly solar energy, has become increasingly critical.11 Photocatalysis has been verified to possess distinct advantages over conventional chemical methods, including operational simplicity, reduced energy demands, and inherent environmental compatibility, owing to its superiority in ambient operation and direct solar-to-chemical energy conversion.12,13 This technology has demonstrated considerable potential in applications such as H2O splitting, CO2 reduction and H2O2 production, offering a promising pathway for energy supply and environmental protection.14 Photocatalytic H2O2 production represents a sustainable innovation, surpassing the energy-intensive multi-step anthraquinone process in industry.15,16 The basic principle of photocatalytic H2O2 production is shown in Fig. 2(a).17


image file: d5mh00697j-f2.tif
Fig. 2 (a) Schematic of the photocatalysis process. CB and VB represent the conduction band and valence band, respectively. (b) Energy diagram for oxygen reduction reaction (ORR) and water oxidation reaction (WOR) on a semiconductor photocatalyst for H2O2 production.

Recently, the photocatalytic synthesis of H2O2 has become a research hotspot. Various novel materials and surface modification strategies have been developed to improve photocatalytic H2O2 production performance, and many exciting progresses have been achieved. On the one hand, studies have demonstrated that the synthesis selectivity and efficiency depend critically on the component and band structure of the applied photocatalyst. The development of new photocatalysts, including TiO2, C3N4, and covalent organic framework (COFs), has significantly enhanced both the H2O2 evolution rate and selectivity.18,19 On the other hand, modifications of particulate photocatalysts have been explored to improve H2O2 generation efficiency, including molecular engineering, defect regulation, and element doping (e.g., nitrogen, sulfur, and phosphorus).20–22 These approaches enhance photocatalytic performance by optimizing light absorption and charge utilization. Previous reviews have mainly focused on photocatalysts based on their preparation methods or reaction pathways.2,4 However, to date, there are no reviews relating to the significant advancements in photocatalytic H2O2 production from the perspective of surface engineering strategies.

The surface of a particulate photocatalyst plays an important role in the transfer and utilization of charge carriers, adsorption and activation of reactant molecules, promotion of H2O2 formation and inhibition of H2O2 decomposition. Various surface engineering strategies have been developed to improve the photocatalytic H2O2 production performance, and many exciting progresses have been achieved recently. However, to date, no specific review has systematically covered this critical topic. To fill this gap, serial surface engineering strategies are classified and elaborated in this review to provide a comprehensive and systematic illustration of this topic. The surface engineering strategies for light-harvesting semiconductors cover surface modification strategies and surface compositing strategies. The former mainly involves crystal facet engineering, surface vacancy engineering and surface functional group regulation to tailor material properties, and the latter includes the integration of functional components, such as metals, semiconductors, carbon nanomaterials, polyoxometalates (POMs) and their derivatives, and organic compounds to enhance specific functionality. Therefore, the challenges and prospects in surface engineering strategies of particulate photocatalysts are also analysed and summarized, which is anticipated to offer valuable guidance for the design and fabrication of advanced particulate photocatalysts for efficient H2O2 synthesis.

2. Basics of photocatalytic H2O2 production

2.1 Oxygen reduction reaction (ORR) to produce H2O2

The ORR is an important step in the photocatalytic production of H2O2, where O2 molecules are reduced to H2O2 by accepting electrons and protons. In the ORR for H2O2 production, two primary mechanisms exist: the two-electron direct reduction pathway and the superoxide radical (˙O2)-mediated pathway. In the two-electron direct reduction pathway, O2 with the participation of a proton accepts two electrons to form H2O2 in one step. This reaction can be expressed as eqn (1):
 
O2 + 2H+ + 2e → H2O2. (1)

In the ˙O2-mediated pathway, electron transfer occurs in two sequential steps. First, O2 captures an electron to form a ˙O2 (eqn (2)). Then, the ˙O2 reacts with another electron and two protons to produce H2O2 (eqn (3)):

 
O2 + e → ˙O2, (2)
 
˙O2 + 2H+ + e → H2O2. (3)

The key distinction between these two mechanisms is that the one-step two-electron pathway is easier to occur thermodynamically and has fewer side reactions.23 In contrast, the two-step single electron pathway is more advantageous kinetically.

2.2 Water oxidation reaction (WOR) to produce H2O2

WOR is another common pathway to H2O2 production, primarily occurring through two mechanisms: the direct two-electron pathway and the hydroxyl radical (˙OH)-mediated pathway. In the direct two-electron pathway, H2O molecules are oxidized, undergoing a two-electron transfer process to form H2O2 (eqn (4)):
 
2H2O → H2O2 + 2H+ + 2e. (4)

However, the ˙OH-mediated pathway involves ˙OH acting as an intermediate. Initially, H2O molecules undergo oxidation, generating highly reactive ˙OH (eqn (5)). Subsequently, two ˙OH couple together to form H2O2 (eqn (6)):

 
H2O → ˙OH + H+ + e (5)
 
2˙OH → H2O2. (6)

Owing to the high reactivity of ˙OH, some side reactions are prone to occur in this route. Although WOR has primarily been investigated for H2O2 production, H2O2 can only become the dominant product over appropriate photocatalysts under controlled reaction conditions. Efficient H2O2 generation requires photocatalysts that can activate H2O molecules and precisely modulate the electron–proton transfer kinetics to stabilize the active intermediates.

2.3 Integrated dual-channel pathway (overall photocatalytic H2O2 evolution reaction)

Through rational design, H2O2 can be co-produced by a dual-channel pathway, including both ORR and WOR. The integration of ORR and WOR establishes a highly efficient H2O2 generation system, wherein ORR consumes electrons and protons, while WOR provides both electrons and protons via H2O oxidation. This electron-proton coupling regulation ensures the atomic utilization of reactants (100%).24 Through rational surface modification, the active sites and surface electronic structure can be optimized to enhance photocatalytic activity and selectivity. Additionally, the band structure should be elaborately tailored to fulfil the energy requirements of the ORR and WOR, as presented in Fig. 2.

2.4 Kinetics of H2O2 production and decomposition

The H2O2 production and decomposition kinetics jointly dominate the net accumulation efficiency of its photocatalytic system, and this dynamic balance directly determines the yield and stability in practical applications. The decomposition of H2O2 in the photocatalytic system mainly involves the following two mechanisms: (1) photocatalytic decomposition under ultraviolet (UV) light, where H2O2 absorbs UV light and decomposes into ˙OH and other reactive oxygen species;25 (2) photocatalytic redox decomposition, where H2O2 interacts with the photogenerated electrons and holes on the surface of the photocatalyst, forming the reactive oxygen species that further accelerate the decomposition of H2O2. The formation and decomposition kinetics of H2O2 can be described using eqn (7):
 
image file: d5mh00697j-t1.tif(7)
where Kf and Kd denote the rate constants of the formation (μmol L−1 min−1) and decomposition (min−1) of H2O2, respectively. A larger Kf and a smaller Kd suggest higher activity toward H2O2 production.26 By optimizing the surface electron density and active sites, and regulating the band structure, the H2O2 decomposition can be effectively suppressed, leading to improved accumulation concentration of H2O2.

2.5 Performance evaluation

The apparent quantum efficiency (AQE) is a key metric for evaluating the efficiency of photocatalytic H2O2 production.27,28 It is defined as the ratio of the number of electrons and holes for producing H2O2 molecules to the number of incident photons at a specific wavelength. The amount of H2O2 produced during the reaction is quantified using chemical analysis methods (e.g., spectrophotometry and iodometry), while the number of incident photons refers to the total number of photons illuminating the photocatalyst surface during the reaction.

The solar-to-chemical energy conversion efficiency (SCC) of photocatalytic H2O2 production is a critical performance metric reflecting the ability of the system to convert solar energy into chemical energy. This efficiency is generally influenced by factors including light absorption, charge separation, and surface reaction kinetics.29 It is typically calculated using the following eqn (8):

 
image file: d5mh00697j-t2.tif(8)
where ΔG represents the change in Gibbs free energy required for H2O2 production, typically 98.2 kJ mol−1. R denotes the rate of H2O2 production per unit time (mol s−1), A is the irradiation area (m2), and P is the solar irradiance per unit area (W m−2). When evaluating and comparing the H2O2 performance, both the AQE and SCC are two key indices.30 These metrics assess the effectiveness and efficiency of photocatalytic H2O2 production from different perspectives, which can provide a more comprehensive evaluation of the photocatalytic efficiency and the practical application potential.

The fundamental photocatalytic H2O2 production process is that the semiconductor is photoexcited to produce electrons and holes that separate and migrate to the photocatalyst surface and participate in redox reactions with the adsorbed reactants. Given that photocatalytic reactions occur on the surface of the photocatalyst, the surface properties and functional components can play an important role in photocatalytic H2O2 production by affecting the charge utilization, reactant adsorption and activation, reaction route, H2O2 accumulation and decomposition. To address these challenges and develop efficient photocatalytic H2O2 production systems, this review focuses on surface engineering strategies based on light harvesting semiconductors in the following two aspects: (1) surface modification strategies, including crystal facet engineering, surface vacancy engineering and functional group tailoring, to precisely modulate the surface properties of semiconductors; (2) surface composite strategies involving the integration of functional components, such as metals, semiconductors, carbon nanomaterials, polyoxometalates (POMs) and their derivatives, as well as organic compounds. A schematic illustration of these surface engineering strategies for particulate photocatalysts for H2O2 production is summarized, as depicted in Fig. 3.


image file: d5mh00697j-f3.tif
Fig. 3 Schematic of typical surface engineering strategies over particulate photocatalysts for promoted H2O2 production.

3. Surface modification strategies

The physicochemical properties of semiconductors vary with different surface structures, which in turn influence their photocatalytic H2O2 production performance by modulating key processes, such as light absorption, charge separation, reactant adsorption and activation, reaction pathways, and product desorption.31 In this section, the surface modification strategies of semiconductors, including crystal facet engineering, surface vacancy engineering, and functional group regulation, are systematically discussed, and their impact on photocatalytic H2O2 production is also discussed afterwards (Table 1).
Table 1 Summary of the surface modification strategies for particulate photocatalysts for H2O2 production
Strategy Photocatalyst Light source H2O2 (μmol h−1 g−1) Reaction solution AQYh (%) Ref.
a MIL-125: Ti-based metal organic framework derived from Ti metal ions and carboxylate organic ligands.b Cv-g-C3N4: g-C3N4 with C vacancies.c DCN-15A: defected g-C3N4 with surface NH2 vacancy.d Bpt-CTF: covalent triazine frameworks with 1,3-bis(4-cyanophenyl)thiourea precursor.e SO3H-COF: covalent organic framework with sulfonic acid groups.f TBTN-COF: cyanide-based covalent organic framework combining 2,4,6-trimethylbenzene-1,3,5-tricarbonitrile.g BM-Au/TiO2: gold nanoparticle-loaded rutile TiO2 with a bimodal size distribution around 10.6 nm and 2.3 nm.h AQY: apparent quantum yield.
Crystal facet engineering TiO2(001) 300 W Xe lamp 649.2 10 vol% isopropanol solution 32
MIL-125a 500 W Xe lamp, λ ≥ 400 nm 0.57 Water 33
BiOCl(001) 300 W Xe lamp 3850 5 vol% formic acid solution 34
Surface vacancy engineering Cv-g-C3N4b 300 W Xe lamp, λ ≥ 420 nm 90 Water 35
DCN-15Ac AM1.5G, λ ≥ 420 nm 83.3 20 vol% isopropanol solution 36
Defect ZrS3 AM 1.5G 2603 30 mL water with 1 mmol benzylamine 37
Surface functional group regulation Bpt-CTFd 300 W Xe lamp 6536.2 Water 8.6 (400 nm) 38
SO3H-COFe 300 W Xe lamp, λ ≥ 400 nm 497.1 10 vol% methanol solution 15 (400 nm) 39
TBTN-COFf 300 W Xe lamp, λ ≥ 420 nm 1376.6 Water 7.6 (420 nm) 40
BM-Au/TiO2g 300 W Xe lamp, λ ≥ 430 nm 35 Water 41


3.1 Crystal facet engineering

The surface structure of a certain crystal is determined by its specific crystal facets exposed, with each crystal facet exhibiting distinct atomic arrangements, electronic structures, and surface energies, which can affect photocatalytic H2O2 production performance in the following issues.42 (1) The arrangement and coordination of surface atoms affect the adsorption of molecules or ions (Fig. 4(a)). (2) The crystal orientations determine the separation of the photogenerated charge carriers (Fig. 4(b)). (3) The redox abilities of charge carriers depend on the surface electronic band structure, which relies on crystal facets (Fig. 4(c)). Therefore, the preparation of semiconductors exposed to suitable crystal facets is an effective way to regulate their photocatalytic H2O2 production performance. Crystal facet engineering is applicable to photocatalysts with anisotropic crystalline structures, where different crystallographic planes exhibit distinct photocatalytic performances. This approach is particularly effective for some oxide semiconductors (e.g., TiO2, ZnO, and BiVO4).
image file: d5mh00697j-f4.tif
Fig. 4 Roles and applications of crystal facet engineering in photocatalytic H2O2 production. Schematics illustrating the important roles of facets: (a) reactant adsorption, (b) charge separation, and (c) redox ability due to the distinct band structure. (d) Free energy diagram of TiO2 with different exposed facets for photocatalytic H2O2 production. Reproduced with permission.32 Copyright 2023, Elsevier. (e) Illustration of the spatial separation of h+/e between (001) and (111) facets on a Ti-based metal organic framework derived from Ti metal ions and carboxylate organic ligands. Reproduced with permission.33 Copyright 2022, the Royal Society of Chemistry. (f) Schematic of the band structure of the BiOCl (001) and BiOCl (010) samples. Reproduced with permission.34 Copyright 2021, the Royal Society of Chemistry.

The adsorption of reactant molecules is highly related to the exposed facet of the semiconductor. For example, Gao and co-workers reported that the anatase TiO2(001) facet played a significant role in photocatalytic H2O2 production. The higher activity of the (001) facet than those of the (101) and (100) facets was attributed to the higher intrinsic activity of the active sites. Theoretical calculations indicated that the under-coordinated Ti atoms on the (001) surface served as strong O2 adsorption sites (Fig. 4(d)). This study provided an insight into the relationship between the crystal facet structure and photocatalytic activity at the atomic scale.32 Besides, semiconductors with different exposed facets show distinct adsorption capacities of surface modifying agents. A large number of hydroxyl groups on the TiO2(001) facet tended to bond with stearic acid (SA) molecules. The enhanced hydrophobicity with a higher adsorbing proportion of SA led to an increased accumulation of O2 in the interfacial microenvironment, thereby enhancing the photocatalytic H2O2 generation rate.43

Charge separation, another significant step in photocatalytic H2O2 evolution, can also be regulated by facet engineering. Pan's group found that meso-tetra (4-carboxyphenyl) porphyrin with a selectively exposed (400) facet could exhibit a higher photocatalytic H2O2 generation rate than that with an exposed (022) or (020) facet. The higher photocatalytic performance was attributed to the strong internal electric field on the (400) facet, which could effectively prohibit charge recombination.44 In addition, the formation of facet junctions can affect the separation of photogenerated electrons and holes between two different facets. The oxidative sites are located on the (001) facet of a Ti-based metal organic framework (MOF) derived from Ti metal ions and carboxylate organic ligands (MIL-125) photocatalyst, while the reduction reaction occurs on its (111) facet. Therefore, such a formed (001)/(111) facet junction of MIL-125 could achieve the spatial separation of redox reaction sites, avoiding the possible recombination of charge carriers (Fig. 4(e)).33

The energy band position of the semiconductor is an important index for determining the reaction pathway, which can be effectively tailored by crystal facet engineering. Xu and co-workers demonstrated that BiOCl with exposed (010) facets produced H2O2 via both the ORR and WOR, while only ORR could occur on BiOCl with the exposed (001) facet (Fig. 4(f)). This is because the conduction band minimum (CBM) of BiOCl with the exposed (001) (−1.28 V vs. NHE) or (010) (−1.03 V vs. NHE) facet is more negative than the oxidation potential of ˙O2/O2 (−0.33 V vs. NHE). However, only holes in the valence band maximum (VBM) of BiOCl with the exposed (010) facet (2.19 V vs. NHE) can oxidize H2O to ˙OH (1.99 V vs. NHE).34 Crystal facet engineering enables the precise control over the redox ability of photocatalysts, thereby influencing the reaction pathway and efficiency of the photocatalytic H2O2 production.

3.2 Surface vacancy engineering

Surface vacancy, as a typical structural defect formed by atom deletion, further enriches the surface chemical environment of the photocatalyst.45 Appropriate surface vacancy regulates the electronic structure of the photocatalyst, thereby affecting light absorption (Fig. 5(a)) and charge separation (Fig. 5(b)). In addition, the surface vacancy participates in the construction of the active site to regulate the reactant adsorption, reaction pathway, and product desorption (Fig. 5(c)). However, excessive vacancies can act as charge recombination centers, resulting in decreased photocatalytic activity.46 Therefore, creating favorable vacancies on the material surface is important to improve the performance of photocatalytic H2O2 production.47 Surface vacancy engineering is widely applied to modulate semiconductors, especially for g-C3N4, metal oxides, perovskite oxides, metal sulfides and oxyhalides. Herein, the roles of surface vacancy in photocatalytic H2O2 production performance are analyzed and discussed to offer an insight into the rational design of photocatalysts, hoping to further deepen the understanding of the relationship between surface chemical environment and photocatalytic performance.
image file: d5mh00697j-f5.tif
Fig. 5 Roles and applications of the surface vacancy of semiconductors for photocatalytic H2O2 production. Schematics of the important roles of surface vacancy in tuning (a) light absorption, (b) charge separation, and (c) reactant adsorption. Density of states of (d) g-C3N4 and (e) defective g-C3N4 with N vacancies. (f) Schematic of mechanisms underlying the photoexcited dynamics involved in photocatalytic H2O2 evolution over g-C3N4 with N vacancies. Reproduced with permission.36 Copyright 2018, Wiley-VCH. (g) The changed H2O2 generation pathway from single-electron reduction on g-C3N4 to two-electron reduction on g-C3N4 with C vacancies. Reproduced with permission.35 Copyright 2016, Elsevier.

The introduced surface vacancy can modulate the electronic structure of photocatalysts, affecting their light absorption ability. Specifically, compared with the conduction and valence band states, weak bonding at vacancy sites reduces the splitting between bonding orbitals and antibonding orbitals, which facilitates the formation of distinct electronic states within the band. There are two primary mechanisms for the enhanced light absorption derived from such electronic states: (1) narrowing the bandgap to enlarge the light absorption wavelength range, and (2) acting as midgap states, which can accommodate the photogenerated electrons from the valence band, thereby enabling the absorption of photons with energies below the intrinsic bandgap threshold.48 For instance, Ye's group introduced surface N vacancies in g-C3N4 (DCN) by a one-step photo-assisted route using a hydrazine as a reducing agent. Density functional theory calculations demonstrate that the DCN showed a narrower bandgap and a defect state within the band (Fig. 5(d) and (e)). The experimental results revealed that DCN exhibited broader and stronger absorption tails extending to 600 nm than g-C3N4. Consequently, the H2O2 production performance of DCN was 10.1 times higher than that of g-C3N4.36 The surface vacancy plays a pivotal role in impacting light absorption to regulate photocatalytic performance by inducing midgap states and narrowing the bandgap.

Besides light absorption, surface vacancies play a crucial role in modulating charge separation. Surface vacancies can act as charge carrier trapping sites to regulate the separation of electrons and holes. For example, Tian and co-workers prepared defective ZrS3 nanobelts with S22− and S2− vacancies and systematically investigated their effects on modulating charge carrier dynamics. Specifically, it was demonstrated that S22− vacancies could prohibit charge carrier recombination, while S2− vacancies could accelerate the extraction of the photogenerated holes towards the surface and promote electron transfer.37 Shi and co-workers constructed defected g-C3N4 with surface NH2 vacancy (DCN-15A) by photothermal treatment. The introduction of an N vacancy in g-C3N4 could lead to the formation of defect states, which facilitated the charge transfer process and prohibited the recombination of electrons and holes (Fig. 5(f)).36

After the photogenerated electrons and holes transfer to the photocatalyst surface, the reaction between the reactants and the charge carriers occurs only when the reactants are effectively adsorbed on the active sites. The coordinatively unsaturated atoms at the vacancy sites promote the adsorption of reactant molecules. Zhang and co-workers induced N vacancies into g-C3N4 (Nv-CN). The calculated adsorption energy revealed that the O2 adsorption energy on Nv-CN was lower than that on g-C3N4, indicating that the surface N vacancy sites were more favorable for the adsorption of O2. The strong O2 desorption peak and increased O–O bond length in Nv-CN further experimentally demonstrated that N vacancies promoted O2 adsorption and activation.19 In addition, surface vacancies play a decisive role in modulating the reaction pathways by enhancing reactant adsorption and stabilizing key intermediates at defect sites. Wang's group observed that the photocatalytic H2O2 activity of g-C3N4 with C vacancies (Cv-g-C3N4) was 14-fold higher than that of the pristine g-C3N4. The experimental results demonstrated that C vacancies changed the ORR pathway from a two-step single-electron indirect reduction to a more efficient one-step two-electron direct reduction (Fig. 5(g)).35 Therefore, the surface vacancy can affect the reactant adsorption and the reaction pathway, thus influencing the final efficiency of photocatalytic H2O2 generation.

3.3 Surface functional group regulation

Surface functional group modification involves the introduction or precise regulation of specific chemical groups on the photocatalyst surface to tailor its physicochemical properties and modulate its photocatalytic performance.49–52 The modification of functional groups can mainly modulate the O2/H2O2 adsorption, charge separation and surface reactions, thereby influencing the photocatalytic H2O2 production performance.53–55 For example, polar functional group modification can affect the charge separation; electron-withdrawing functional group modification can regulate both processes of charge separation and O2 adsorption; surface protonation provides a proton-rich environment for ORR; surface complexation or adsorption of cationic/anionic species can mediate the H2O2 adsorption and decomposition (Fig. 6(a)). The surface functional group modification strategy is effective for materials featuring abundant surface-active groups (e.g., –OH and –NH2) or those amenable to directional chemical functionalization, such as g-C3N4, MOFs, COFs, and TiO2.
image file: d5mh00697j-f6.tif
Fig. 6 Roles and applications of the surface functional group modification in photocatalytic H2O2 production. (a) Schematics illustrating the roles of surface functional group modification in photocatalytic H2O2 production. (b) Rational synthesis of Bpu-CTF and Bpt-CTF photocatalysts featuring different charge separation and transfer abilities. The CTF represents covalent triazine frameworks; the Bpu represents 1,3-bis(4-cyanophenyl)urea precursor; and the Bpt represents 1,3-bis(4-cyanophenyl)thiourea precursor. Reproduced with permission.38 Copyright 2022, Wiley-VCH. (c) Mechanism for photocatalytic H2O2 production on the surface of cyanide-based COF combining 2,4,6-trimethylbenzene-1,3,5-tricarbonitrile. Reproduced with permission.40 Copyright 2024, Wiley-VCH. (d) A proposed mechanism of the photocatalytic ORR for H2O2 production based on Au/TiO2-CO32−. Reproduced with permission.41 Copyright 2016, Wiley-VCH.

Polar functional groups create a built-in electric field (BEF) that enhances charge separation. The BEF forms dipoles that drive electrons and holes in opposite directions. Han's group introduced polar thiourea moieties onto covalent triazine frameworks (CTF) to construct the BEF, enhancing the charge separation (Fig. 6(b)).38 Then, a series of triazine COF photocatalysts functionalized with oxygen groups of varying polarity was further prepared and analyzed. The results showed that the strongly polar sulfonic acid groups (–SO3H) generated powerful BEF, boosting charge separation and H2O2 production.39 The modification with electron-withdrawing functional groups can play dual functions in regulating both charge separation and O2 adsorption. A representative example is the cyano group (–CN) that has been extensively employed for photocatalytic H2O2 production.56 Zhou and co-workers designed a cyanide-based COF combining 2,4,6-trimethylbenzene-1,3,5-tricarbonitrile (TBTN-COF). The TBTN-COF exhibited more efficient charge separation and higher charge carrier availability than its counterpart without –CN. Besides, the –CN sites with concentrated electron density acted as O2 adsorption sites, facilitating the formation of *OOH intermediates and a favorable ORR pathway (Fig. 6(c)).40 In addition to –CN, the electron-withdrawing groups, such as sulfone group, hydroxyl group, vinyl group and benzothiadiazole units, have also been applied to regulate the charge separation and O2 adsorption during photocatalytic H2O2 evolution.57–60

Proton is another functional group for surface modification.61–63 The protonated surface serves as a proton enrichment platform that can directly supply protons for ORR. The abundant proton facilitates the establishment of a new reaction equilibrium by accelerating the forward reaction while inhibiting the reverse reaction. Wen and co-workers reported a protonation and an O-doped g-C3N4 photocatalyst. The abundant protons on the surface can combine with the adsorbed O2, thus promoting the formation of the OOH* intermediate and facilitating the H2O2 production.64 Surface protonation is an effective strategy to provide protons for facilitating the H2O2 formation, avoiding the usage of an acid solution for the reaction.

In addition, the complexation and adsorption of ions have emerged as effective approaches to regulating the process of photocatalytic H2O2 production. The H2O2 decomposition occurs on the TiO2 surface because H2O2 tends to complex with the surface Ti–OH group to form the Ti–OOH, which can be reduced by the photogenerated electrons. Maurino and co-workers found that photocatalytic H2O2 production over TiO2 was undetectable; however, it reached a stable millimolar concentration after adding F. They proposed that the competitive adsorption of F and H2O2 on the TiO2 surface hindered the H2O2 adsorption and subsequent decomposition.65 Teranishi and co-workers found that CO32− ions chemisorbed on the TiO2 surface could inhibit the H2O2 decomposition by suppressing a Ti–OH-mediated degradation pathway (Fig. 6(d)).41 By leveraging the interaction between photocatalyst surfaces and ions, the H2O2 decomposition can be finely regulated. The versatility of surface complexation and adsorption of cationic/anionic species highlights their potential for designing highly efficient and stable photocatalytic H2O2 production systems.

In this section, surface modification strategies, including crystal facet engineering, surface vacancy engineering, and functional group regulation, enable atomic/molecular-level regulation of the physicochemical properties of semiconductors, thereby optimizing the key steps in photocatalytic H2O2 synthesis. Crystal facet engineering and surface vacancy engineering can promote intrinsic activity through atomic-level structural optimization, but their applications are constrained by limited material varieties and challenging synthesis approaches. Crystal facet engineering is mainly applicable to inorganic semiconductors with anisotropic crystalline structures. However, the modulation of the surface vacancy depends highly on the controllability of the surface structure. The electronic structure and chemical microenvironment can be easily tuned by grafting the photocatalyst surface with specific functional groups. Functional group modification is mainly applicable to organic semiconductors, which offer greater diversity and flexibility but suffer from instability under specific conditions. Therefore, continuous exploration of surface modification strategies is necessary to develop efficient and stable photocatalysts. In addition, these strategies can be combined with surface composite strategies, as discussed in later sections, to collectively improve the H2O2 production efficiency.

4. Surface composite strategies

Surface composite strategies are another part of effective surface engineering, whose emphasis lies in the integration with other materials to achieve the function-oriented optimization of photocatalysts.2,4,66,67 Distinct from surface modification approaches that focus on altering the intrinsic surface properties of photocatalysts, surface composite strategies can benefit from the advantages of both materials, thereby effectively alleviating the inherent limitations of photocatalyst systems with only a single material. In this section, the surface composite strategies are overviewed (Table 2).
Table 2 Summary of the surface composite strategies for particulate photocatalysts for H2O2 production
Strategy Photocatalyst Light source H2O2 (μmol h−1 g−1) Reaction solution AQYk (%) Ref.
a C-W18O49/Ni-SA: Ni single-atom-decorated C-W18O49 photocatalyst.b CoSA/Py-CTF: Co single-atom-modified CTF photocatalyst.c Sb-SAPC: Sb single-atom decorated polymeric carbon nitride.d CN/Zn-MOF(lc): composite photocatalyst with g-C3N4 and Zn-based metal organic framework.e CuO/Cu(OH)2@CN: composite photocatalyst of CuO, Cu(OH)2 and g-C3N4.f α-Fe2O3/CQD@CN: composite photocatalyst of α-Fe2O3, carbon quantum dots and g-C3N4.g CQD-CTF: composite photocatalyst of carbon quantum dots and covalent triazine frameworks.h g-C3N4-PW11: composite photocatalyst of g-C3N4 and polyoxometalate cluster of [PW11O39]7−.i g-C3N4-SiW11: composite photocatalyst of g-C3N4 and polyoxometalate cluster of [SiW11O39]8−.j g-C3N4/BDI: electron-deficient biphenyl diimide units modified g-C3N4 photocatalyst.k AQY: apparent quantum yield.
Composite with metals C-W18O49/Ni-SAa 300 W Xe lamp, λ ≥ 420 nm 338.9 Water 28.51 (365 nm) 68
CoSA/Py-CTFb 300 W Xe lamp, λ ≥ 420 nm 1159.32 Water 13.2 (420 nm) 69
TiO2/MoSx–Au 300 W Xe lamp 3044 10 vol% ethanol solution 7.2 (365 nm) 70
AgPd/BiVO4/CoOx LED lamp, λ ≥ 400 nm 1560 Water 5.8 (420 nm) 71
Sb-SAPCc 300 W Xe lamp, λ ≥ 420 nm 647 Water 17.6 (420 nm) 6
CN/Zn-MOF(lc)d 300 W Xe lamp, λ ≥ 400 nm 1912 20 vol% ethanol solution 72
Composite with semiconductors ZnO/WO3 300 W Xe lamp 3.364 Water 0.17 (365 nm) 73
CuO/Cu(OH)2/CNe 300 W Xe lamp, λ ≥ 420 nm 677 Water 74
In2S3/CdS AM 1.5G 836 Water 75
Composite with carbon nanomaterials Fe2O3/CQD@CNf 300 W Xe lamp, λ ≥ 420 nm 69.3 Water 17.89 (420 nm) 76
CQD-CTFg 300 W Xe lamp, λ ≥ 420 nm 259 Water 1.03 (420 nm) 77
rGO/TiO2 300 W Xe lamp, λ ≥ 320 nm 1500 5 vol% isopropanol solution 78
Composite with POMs or their derivatives g-C3N4-PW11h 300 W Xe lamp 3.5 Water 79
g-C3N4-SiW11i AM 1.5G 178 5 vol% ethanol solution 6.5 (420 nm) 80
Composite with organic compounds g-C3N4/BDIj AM 1.5G, λ ≥ 420 nm 28.3 Water 4.6 (420 nm) 81
TiO2 AM 1.5G, λ ≥ 300 nm 490 Furfural alcohol solution 82


4.1 Composite with metals

Metals, usually employed as reduction cocatalysts, play a pivotal role in modulating the electronic structure of photocatalysts, influencing the activity and selectivity of photocatalytic H2O2 production. Researchers have developed various metal cocatalysts for photocatalytic H2O2 production via ORR, but the advancement of cocatalysts targeting WOR remains limited. Generally, noble metals, such as Au, Pd, Pt, and Ag, are commonly utilized as cocatalysts for photocatalytic H2O2 production via the route of ORR.83–85 Considering the high cost of noble metals, alternatives based on non-noble metals, including Ni and Co, have also been explored. Metal cocatalyst modification is a widely applicable surface engineering strategy. However, its effective implementation crucially depends on lattice matching, electronic structure compatibility, and proper band alignment between the cocatalyst and the semiconductor.
4.1.1 Composite with metal nanoparticles. Metal nanoparticles as an efficient ORR cocatalyst can significantly accelerate the reaction kinetics process owing to their unique electronic structure. Based on their particle size, metal cocatalysts can be categorized into metal nanoparticles, metal clusters, and metal single atoms. When co-catalysts are deposited on the semiconductor, the efficiency of ORR is generally improved mainly by enhancing the electron separation and transfer, and lowering the activation energy or the overpotential of ORR. What is more, metal cocatalysts with distinct electronic and geometric configurations exhibit different affinities for adsorbates, resulting in distinct O2 adsorption behaviors and H2O2 desorption processes, which determine the photocatalytic activity and selectivity. Generally speaking, the photocatalytic H2O2 performance can be significantly enhanced by precisely regulating the species, size, morphology, and dispersion of metal nanoparticle cocatalysts.86,87

Au, Pd, Pt, Ag, and Ni nanoparticles have commonly been used as effective ORR co-catalysts for photocatalytic H2O2 production, among which Au nanoparticles are the most widely studied.88 However, the weak O2 adsorption ability on the Au cocatalyst, caused by its intrinsic electronic structure, limits its photocatalytic performance of H2O2 production. According to molecular orbital theory, the adsorption energy of the adsorbate is determined by the antibonding-orbital occupancy degree between the metal sites and adsorbates. Yu's group regulated the electronic structure of Au nanoparticles by introducing MoSx as an electron mediator to optimize the O2 adsorption.70 For the TiO2/MoSx–Au photocatalyst, Au nanoparticles were selectively deposited onto MoSx, which was subsequently anchored on TiO2. The electron-deficient Auδ+ active sites emerged after introducing MoSx, stemming from directional electron transfer from Au nanoparticles to MoSx, as shown in Fig. 7(a). This electronic redistribution suppressed the antibonding orbital occupancy of Au–Oads and enhanced O2 adsorption. Pd is another typical cocatalyst for photocatalytic H2O2 production, showing a stronger O2 affinity than Au owing to the symmetry-adapted linear combination of Pd 4d orbitals and O 2p orbitals. Chu's group had precisely loaded Pd and CoOx cocatalysts on the {010} and {110} facets of BiVO4:Mo, respectively, using a photodeposition method.89 The Pd cocatalyst served as an active site for O2 adsorption and steered two-electron ORR for H2O2 production. Further experimental results proved that the selective deposition of Pd and CoOx cocatalysts tailored the energetics of {010} and {110} facets, respectively, thus promoting charge separation and suppressing charge carrier recombination. Moreover, Xu's group loaded a series of metal nanoparticles onto the WO3 photocatalyst and investigated their effects on the photocatalytic H2O2 production.90 They found that the order of kf values was Pd > Au > Pt > Ag, while the order of kd values was Pt > Pd > Ag > Au. These results revealed the distinct photocatalytic characteristics of different metal cocatalysts.


image file: d5mh00697j-f7.tif
Fig. 7 Application of particulate photocatalysts with metal cocatalysts for photocatalytic H2O2 production. (a) Schematic of electron-deficient Auδ+ formation to reinforce Au–Oads bond: (1) weak Au–Oads bond on the Au surface, (2) the formation of electron-deficient Auδ+ sites via MoSx incorporation, and (3) strong Au–Oads bond on MoSx–Au surface. Reproduced with permission.70 Copyright 2024, Nature Publishing Group. Energy-band diagrams for (b1) the Au/TiO2, (b2) Ag/TiO2, and (b3) AuAg/TiO2 heterojunctions. Evac, EF, ΦM, Φb, and χ denote the vacuum level, Fermi level, work function of metal, Schottky barrier height, and electron affinity of the TiO2 conduction band, respectively. (b4) Mechanism for the photocatalytic production of H2O2 on the AuAg/TiO2 photocatalyst. Reproduced with permission.13 Copyright 2012, American Chemical Society. (c) Schematic of the charge separation process on CoOx/BiVO4/(Ag/Pd). Reproduced with permission.71 Copyright 2022, Nature Publishing Group. (d) End-on O2 adsorption type on an isolated atomic site. Reproduced with permission.6 Copyright 2021, Springer Nature Limited.

However, the photocatalytic H2O2 performance is only moderately improved for the photocatalysts with monometallic cocatalysts owing to the following aspects. First, the electronic structures of monometallic cocatalysts often mismatch with the diverse energy levels of photocatalysts, and high Schottky barriers impede electron transfer. Second, the current cocatalysts usually exhibit an electron configuration mismatch between the adsorbed O2 and the metal sites, leading to either excessively strong or weak in O2 adsorption. Third, some cocatalysts adsorb O2 in a side-on O2 adsorption configuration rather than the optimum end-on O2 adsorption model, resulting in a broken O–O bond and decreased H2O2 selectivity. Alloying offers a promising approach to precisely modulating the electronic and geometric configurations, thereby regulating the photocatalytic H2O2 production process. Tsukamoto's group developed a bimetallic AuAg alloy cocatalyst for efficient photocatalytic H2O2 generation.13 On one hand, the alloying effect significantly promoted the ORR for H2O2 formation. As illustrated in Fig. 7(b1)–(b4), the optimal Schottky barrier height of the AuAg alloy facilitated the efficient electron transfer from TiO2 to the AuAg alloy. Besides, the electron-rich Au sites were created to accelerate the ORR owing to the electronegativity difference between Au and Ag. On the other hand, this alloying strategy effectively suppressed the H2O2 decomposition. The construction of similar core–shell structured cocatalysts is another promising strategy for finely modulating the electronic configurations of monometallic cocatalysts. Chu's group constructed Ag/Pd core/shell cocatalyst to regulate the Schottky barrier between BiVO4 and Pd, which is a major hurdle of the charge separation.71 The Ag core with a low work function served to reduce the Schottky barrier between the {010} facet of BiVO4, while the Pd shell with a high work function could act as the active sites to drive the two-electron ORR (Fig. 7(c)). Overall, the interaction between two metals in the cocatalyst, whether in an alloy or core–shell configuration, can effectively regulate the charge transfer dynamics and O2 adsorption ability by tailoring both electronic and geometric structures, thus affecting the final H2O2 evolution performance.

4.1.2 Composites with metal clusters. Metal cluster cocatalysts, composed of several to hundreds of atoms with sub-nanometer dimensions, can exhibit higher atomic utilization than the case of metal nanoparticle cocatalysts. The local coordination environment of the metal cluster cocatalyst greatly affects its activity and stability. For example, Guo's group confined Pd clusters to fluorinated COFs for photocatalytic H2O2 production.91 The introduction of fluorine not only lowered the d-band center but also strengthened the metal-support interaction between Pd clusters and COFs. Specifically, the lower d-band center weakened the binding energy between the intermediate oxygen species and Pd, facilitating H2O2 desorption from the Pd sites. Improved stability was also observed for the photocatalyst system mainly owing to the enhanced metal-support interaction between Pd clusters and COFs.
4.1.3 Composite with single-atom metals. Unlike a nanoparticle and cluster, a single atom can achieve 100% atomic utilization efficiency and provide a well-defined active site. Therefore, a diverse range of single-atom metal cocatalysts, including noble metals, transition metals, and main group metals, has been extensively investigated for photocatalytic H2O2 production.92–94

The adsorbed O2 molecules on the metal surface can generally be categorized into three types: Pauling-type (end-on), Griffiths-type (side-on), and Yeager-type (side-on). The end-on O2 adsorption configuration minimizes O–O bond breaking, thereby promoting high selectivity for the two-electron ORR. The isolated single-atom site facilitates O2 adsorption in an end-on type, thereby inhibiting O–O bond breaking and improving the activity and selectivity of the two-electron ORR. Ohno's group designed an Sb single-atom decorated polymeric carbon nitride (Sb/PCN) photocatalyst for H2O2 production in pure water under an O2 atmosphere.6 The isolated Sb atomic sites adsorbed O2 in the end-on type, which could inhibit the O–O bond breaking and achieve a favorable two-electron ORR for H2O2 production (Fig. 7(d)).

Besides, the electronic structure of single-atom cocatalysts can be easily regulated by modifying their physicochemical environment. Li and co-workers designed a PCN photocatalyst modified with a Zn single atom.95 The coordination environment of Zn sites could be tailored by changing Zn salt anions. The O atoms in the acetate anion and the N atoms in g-C3N4 coordinated with Zn single-atoms, forming a Zn–N3O moiety. The favorable electronic structure of the asymmetric Zn–N3O moiety promoted O2 adsorption and facilitated the formation of the *OOH intermediate, thus achieving a high photocatalytic H2O2 evolution performance. The complex and diverse coordination structures between single-atom species and supporting materials enable the precise modulation of d-band centers. Jin and co-workers synthesized Ni single-atom-decorated C-W18O49 photocatalyst (C-W18O49/Ni-SA) using an impregnation–calcination method.96 The downshifted d-band center of Ni single-atom led to moderate O2 adsorption, lowered reaction energy barrier for *OOH conversion, and favorable H2O2 desorption. Similarly, Yamashita's group developed a hafnium-based MOF with missing-linker defects and Ni single-atoms to enhance photocatalytic H2O2 production from H2O and O2.72 Ni single-atoms suppressed charge recombination, and the linker defects inhibited H2O2 decomposition. This cooperative effect enabled the H2O2 production activity to be 6.3 times higher than that of the pristine MOF photocatalyst.

Moreover, single-atom cocatalysts provide an ideal platform for identifying the active sites and exploring the structure–performance relationship. Li and co-workers developed a hybrid CN/Zn-MOF photocatalyst with Zn–O4 sites, which is composed of a Zn single atom and four coordinated O atoms.72 The activated Zn single-atom sites facilitated the capture of photoelectrons and the formation of the *OOH intermediate. Zhu and co-workers synthesized a Co single-atom modified CTF photocatalyst (CoSA/Py-CTF) using a ligand engineering strategy.69 The Co single-atoms dispersed on Py-CTF substrates were predominantly in the Co–N3 configuration. The experimental results revealed the dynamic reconstruction behavior of the Co single-atom active sites. Specifically, Co single atoms are dynamically detached from the Py-CTF interface following O2 adsorption under light irradiation. These released atoms were combined with their adjacent counterparts to form transient Co atom pairs, which served as the actual active sites for ORR. This newly formed active site configuration facilitated a transition in O2 adsorption from the Pauling-type to the Yeager-type geometry and enhanced electron transfer from Py-CTF to O2. In summary, single-atom cocatalysts serve as an ideal platform for elucidating the dynamic evolution of active sites and uncovering the structure–performance relationship. The unique geometric structure and distinctive electronic properties of single-atom cocatalysts enable the achievement of specific O2 adsorption configurations, which can facilitate the O2 adsorption, intermediate evolution, and H2O2 desorption.

4.2 Composites with semiconductors

The surface modification of semiconductors by compositing with another semiconductor to form a heterojunction is a powerful strategy that can effectively regulate light absorbance and charge separation. It can be classified into several primary types based on the alignment of the energy levels of the two semiconductors (Fig. 8(a)–(c)), which are type I, type II, and type III. In the type I heterojunction, the band alignment is straddling, where the CBM and VBM levels of the wider bandgap semiconductor are more negative and more positive, respectively, than those of the narrower bandgap semiconductor. In this band alignment, charge carriers tend to accumulate in the narrower bandgap semiconductor, resulting in weaker redox ability and no improvement in charge separation. In the type III heterojunction, the VBM level of one semiconductor lies above the CBM level of the other semiconductor, resulting in no charge transfer between these two semiconductors. Both type I and type III heterojunctions show unfavorable charge transfer characteristics. In contrast, the type II heterojunction features a staggered band alignment in which the CBM and VBM levels of the two semiconductors are offset. This configuration enables the photogenerated electrons and holes to migrate to two different semiconductors, making the spatial separation of charge carriers feasible.97,98
image file: d5mh00697j-f8.tif
Fig. 8 Mechanisms of typical heterojunctions and their application for photocatalytic H2O2 production. Schematic energy diagrams of three different types of heterojunctions: (a) type I, (b) type II, and (c) type III. (d) Photocatalytic production of CO and H2O2 over NH2-MIL-125-Ti/WO3–x heterojunction photocatalyst. Reproduced with permission.73 Copyright 2023, Elsevier. (e) The mechanism of photocatalytic H2O2 production over the CuO/Cu(OH)2@g-C3N4 composite. Reproduced with permission.74 Copyright 2024, Elsevier.

Besides those different types of heterojunctions, another different heterostructure of the Z-scheme has recently been extensively investigated. The original idea was first proposed in 1979 but was experimentally demonstrated in 2001 with the application of the photocatalytic overall water splitting reaction.99,100 Z-scheme systems can be classified into several types according to the difference in mediators and system configurations: aqueous redox mediator in a powder suspension system, solid-state electron mediator in a powder suspension system, and solid-state electron mediator in a sheet system. It is noteworthy that two photocatalysts with well-matched band structures and surface configurations may form a Z-scheme mechanism directly, driving the overall photocatalytic reaction even without the electron mediator. The photocatalytic H2O2 production using a Z-scheme heterojunction photocatalyst is currently conducted following the above approach. In this case, the reduction reaction tends to occur on the semiconductor with a more negative CBM, whereas the oxidation reaction occurs on the other. Then, the remaining holes and electrons in the two semiconductors recombine via the interparticle physical contact or a solid-state electron mediator to complete the entire reaction cycle. Efficient charge recombination at the interface is achieved to enable more photogenerated charge carriers to participate in the corresponding redox reactions.101 Herein, the Z-scheme heterojunction without a solid-state electron mediator is similar to an S-scheme heterojunction.102,103

Enhanced charge separation efficiency can be achieved by constructing effective heterojunction structures.104,105 He and co-workers constructed a heterojunction photocatalyst composed of organic amine-constrained ions intercalated with carbon nitride and CdSe-diethylenetriamine (K+/I-CN/CdSe-D) for H2O2 production.75 The formation of C–Se bonds at the interface between K+/I-CN and CdSe-D significantly improved the electron conductivity and enhanced the charge carrier transfer efficiency. Jiang and co-workers developed an NH2-MIL-125-Ti/WO3–x heterojunction photocatalyst to promote charge separation and transfer by regulating the intensity of the interfacial electric field, which could be increased by maximizing the Fermi level difference between WO3–x and NH2-MIL-125-Ti (Fig. 8(d)).73 In addition, heterojunction photocatalysts demonstrate unique advantages in enabling dual-channel H2O2 synthesis. Generally, it is challenging for a single semiconductor to achieve dual-channel photocatalytic H2O2 production since rigorous redox ability is thermodynamically needed. It is possible to promote charge separation while preserving the strong redox abilities of semiconductors by constructing heterojunctions. For example, Qin and co-workers designed a CuO/Cu(OH)2@g-C3N4 heterojunction photocatalyst for dual-channel H2O2 production.74 In this system, g-C3N4 with the strong reduction ability enabled two-step one-electron ORR, while CuO with sufficient oxidation ability could drive the WOR (Fig. 8(e)).

4.3 Composite with carbon nanomaterials

Carbon nanomaterials formed by sp2 or sp3 hybridization of carbon atoms are commonly used for compositing owing to their unique advantages, including high specific surface area, excellent electrical conductivity, tunable electronic structures, and remarkable chemical stability.106 Therefore, various carbon nanomaterials, including carbon quantum dots (CQDs), carbon nanotubes (CNTs), and reduced graphene oxide (rGO), have been incorporated. Specifically, the integration of carbon nanomaterials with semiconductors offers a powerful strategy to extend light absorption, improve charge transfer, and promote the O2 adsorption, leading to a boosted photocatalytic H2O2 evolution rate. Chen and co-workers developed a ternary photocatalyst of α-Fe2O3/CQD@g-C3N4 for H2O2 production.76 The up-conversion fluorescence spectra revealed that the CQDs featured the highest fluorescence emission intensity at approximately 445 nm, while the light absorption extended into the near-infrared region on account of the up-conversion of the loaded CQDs. In addition, the enhanced charge separation and transfer can also be achieved by incorporating carbon nanomaterials owing to their high conductivity and exceptional electron capture and storage abilities. For instance, Yang and co-workers designed a CTF photocatalyst with CQDs embedded within its porous structure as an electron transport medium, leading to promoted charge separation and transfer efficiency.77 Interestingly, carbon nanomaterials can also act as active sites for the two-electron ORR with high activity and selectivity. Choi's group designed a rGO/TiO2 photocatalyst and demonstrated that rGO could act as an ORR cocatalyst.78 Beyond boosting photocatalytic H2O2 production, carbon nanomaterials play a pivotal role in reinforcing the structural and chemical stability of photocatalysts. For example, Zhu's group prepared rGO-modified black phosphorus quantum dots (rGO@BPQDs) and anchored them on g-C3N4 via π–π interactions and self-trapping pore confinement effect.107 The oxygen-bridged rGO encapsulation created a robust core–shell architecture that effectively suppressed BPQD degradation, enhancing both thermal stability and oxidation resistance by 3.2-fold compared to bare BPQDs. These findings establish carbon nanomaterials as indispensable stabilizers for designing composite photocatalysts that combine high H2O2 productivity with long-term operational durability.

4.4 Composites with POMs or their derivatives

POMs are metal-oxygen cluster compounds composed of oxygen and pre-transition metal ions (such as V, Mo, and W) in high oxidation states. POMs have abundant active sites owing to the abundant metal centers within their well-defined structures. Moreover, POMs can regulate the band structure and electron transfer efficiency of photocatalysts. Zhao and co-workers developed a novel g-C3N4-PW11 hybrid photocatalyst by combining g-C3N4 and a polyoxometalate cluster of [PW11O39]7− (PW11) with an organic linker.79 The covalent combination of PW11 and g-C3N4 could positively shift the VBM level, which was favorable for WOR. In addition, the g-C3N4-PW11 exhibited enhanced separation of charge carriers. They further prepared a hybrid photocatalyst of g-C3N4-SiW11 through the covalent combination of g-C3N4 and the polyoxometalate cluster of [SiW11O39]8− (SiW11).80 The introduction of SiW11 enhanced the O2 adsorption ability, which could facilitate the two-electron ORR for H2O2 production. In addition, considering that the POMs derived metal oxides can accept, transport and store electrons, they constructed a hybrid photocatalyst of g-C3N4 modified with POM-derived metal oxides by calcinating the precursors of g-C3N4 and POMs.108 The enhanced photocatalytic H2O2 performance of g-C3N4-PWO was mainly attributed to the enhanced electron transfer efficiency and improved O2 adsorption capacity. Therefore, it can be concluded that owing to the distinct structural and functional versatility of POMs and their derivatives, the corresponding composite photocatalysts can be further regulated to promote photocatalytic H2O2 production.

4.5 Composites with organic compounds

The primary advantage of using organic compounds to modify semiconductors lies in their versatility and flexibility owing to the diversity of organic compounds and available modification methods, such as chemical bonding, self-assembly, and physical adsorption. Thus, the photocatalytic H2O2 performance can be effectively regulated in terms of charge separation, reactant adsorption and activation. For example, Hirai's group designed an electron-deficient biphenyl diimide (BDI) unit-modified g-C3N4 photocatalyst and investigated its mechanism of photocatalytic H2O2 production in the presence of pure water and O2.81 Experimental and theoretical calculation results revealed that the photogenerated electrons were localized on the melem units while holes were localized on the BDI units. Such spatial charge separation could effectively prohibit the recombination of charge carriers, leading to enhanced H2O2 production. Song's group designed an SA-modified TiO2 photocatalyst for H2O2 evolution (Fig. 9(a)).43 The SA modification increased the O2 accessibility in the interfacial microenvironment. Organic compounds adsorbed or complexed on the photocatalyst's surface can also influence the photocatalytic H2O2 production. For example, Long's group investigated the effect of alcohol complexes on the photocatalyst surface for H2O2 production.82 The surface complexation of benzyl alcohol (BA) or furfural alcohol (FFA) on TiO2 demonstrated the dual functionality of enhancing H2O2 evolution and inhibiting H2O2 decomposition (Fig. 9(b)). Taking FFA as an example, the strong interaction between TiO2 surface and FFA induces a negative shift in the flat band potential, which is favorable for ORR. Besides, the formation of stable FFA complexes suppresses the H2O2 adsorption and subsequent decomposition. These synergistic effects collectively promote the H2O2 production efficiency.
image file: d5mh00697j-f9.tif
Fig. 9 Surface composite of particulate photocatalysts with organic compounds for photocatalytic H2O2 production. (a) Scheme of H2O2 production on the stearic acid modified TiO2 photocatalyst. Reproduced with permission.43 Copyright 2023, American Chemical Society. (b) A two-electron ORR pathway for H2O2 formation on TiO2 in the presence of benzyl alcohol (BA) and furfural alcohol (FFA). Reproduced with permission.82 Copyright 2020, Elsevier.

In this section, surface composite strategies, including the integration of light-absorbing semiconductors with metals, other semiconductors, carbon materials, POMs, and organic compounds, can significantly enhance photocatalytic H2O2 production performance. Semiconductors with insufficient surface active sites can be composited with metals or POMs to enhance intrinsic activity, yet the development of metals and POMs specifically for H2O2 synthesis remains limited. Semiconductors with poor charge transfer characteristics can be effectively alleviated by compositing with carbon nanomaterials or other semiconductors. Moreover, to achieve efficient dual-channel H2O2 synthesis, compositing with other semiconductors to construct a well-designed heterojunction structure is essential. Surface composition strategies should also be a key focus in future studies to facilitate the development of high-performance photocatalytic systems.

5. Summary and outlook

Based on the surface engineering strategies of particulate semiconductor photocatalysts, the recent progress in surface modification and composite strategies to achieve efficient and stable photocatalytic H2O2 production has been systematically summarized in this review, which is anticipated to offer valuable guidance for the design and fabrication of advanced photocatalysts for efficient solar H2O2 synthesis. To date, significant progress has been achieved in photocatalytic H2O2 production, especially in terms of the design and construction of efficient particulate photocatalysts with tailored surface modifications. However, the study of photocatalytic H2O2 production is still in the laboratory mainly because of the challenges of charge utilization and H2O2 decomposition. To further address these great challenges, the surface engineering strategies described here can provide a probable approach that is expected to focus on the following aspects.

(1) It is essential to deepen the understanding of reaction mechanisms and the structure–performance relationship, which can provide a theoretical basis for modifying particulate photocatalysts to promote H2O2 production. Herein, advanced in situ characterization techniques and theoretical calculations are of vital importance to clarify the process mechanisms. Advanced in situ characterization techniques, such as in situ environmental transmission electron microscope, operando X-ray absorption fine structure measurements, and in situ Fourier transform infrared spectroscopy, combined with theoretical calculations can reveal microstructural evolution, interfacial reactions, and intermediate dynamics.

(2) It is necessary to rapidly screen and fabricate complicated surface modifiers over particulate photocatalysts to achieve efficient H2O2 production. Traditionally, finding the optimal combinations of photocatalyst and surface engineering strategies through trial-and-error testing of numerous possibilities is time-consuming. Herein, emerging computational and data-driven tools for photocatalyst design are highly suggested. Machine learning revolutionizes photocatalyst design through data-driven predictive modeling, which uncovers structure–performance relationships across overall chemical spaces and is expected to achieve optimal surface properties of the photocatalyst. Data-driven high-throughput screening using reliable descriptors offers a faster alternative for screening the optimal combinations of photocatalyst and surface engineering strategies and identifying high-performance systems. Large language models can autonomously design, manage, and execute multi-step experiments for developing efficient H2O2 photocatalysts. Integrating the emerging computational and data-driven tools into lab workflows shows a strong potential to perform complex tasks with minimal human input, promising to revolutionize research efficiency and innovation.

(3) It is highly expected to achieve the continuous synthesis of high-concentration and high-purity H2O2 based on modified particulate photocatalysts. The significantly boosted H2O2 formation and remarkably prohibited H2O2 decomposition should be ensured to achieve high-concentration H2O2 synthesis, in which both the rational surface engineering of photocatalysts and the well-designed reaction system need to be developed. It is essential to avoid the use of sacrificial reagents and other additives through surface strategies to produce a high-purity H2O2 solution. To enhance the long-term stability and reusability of photocatalysts, it is necessary to immobilize particulate photocatalysts on the sheet substrate, in which effective surface strategies are essential to construct and stabilize the active sites. Furthermore, developing cost-effective surface engineering strategies with consideration of synthesis complexity and scalability for large-scale production is necessary for future real-world applications.

Photocatalytic H2O2 evolution using particulate photocatalysts has garnered significant attention as a scalable and cost-effective approach to sustainable H2O2 production. In recent decades, remarkable fundamental research progress in this field has been made, achieving a maximum H2O2 concentration of tens of millimoles per liter. This concentration is close to the level of the practical application; for example, a concentration of 1–3 wt% (294–890 mM) is used in H2O2 disinfectants. Therefore, it is anticipated that photocatalytic H2O2 evolution can be developed for future industrial applications. It is our sincere hope that this review will attract more researchers to boost this field and accelerate its development progress in both fundamental and application aspects.

Author contributions

Y. Z. and Z. C. conceived the review and wrote the manuscript. M. L., S. Y. and S. C. conceived the review, reviewed the manuscript, and revised it. All the authors have given approval to the final version of the article and have agreed to be accountable for all aspects of the work in ensuring that questions related to the accuracy or integrity of any part of the work are appropriately investigated and resolved.

Conflicts of interest

There are no conflicts to declare.

Data availability

No primary research results, software or code has been included and no new data were generated or analysed apart of this review.

Acknowledgements

This work was financially supported by the National Natural Science Foundation of China (22272082, 22372001), the Funds for International Cooperation and Exchange of the National Natural Science Foundation of China (W2421036), Anhui Natural Science Foundation for Outstanding Young Scholars (2408085Y008), the Fundamental Research Funds for the Central Universities, Nankai University (63213098) and Starting Fund for Scientific Research of High-Level Talents, Anhui Agricultural University (rc382108).

Notes and references

  1. P. Liu, T. Liang, Y. Li, Z. Zhang, Z. Li, J. Bian and L. Jing, Nat. Commun., 2024, 15, 7823 CrossRef PubMed.
  2. X. Zeng, Y. Liu, X. Hu and X. Zhang, Green Chem., 2021, 23, 1466–1494 RSC.
  3. X. Hu, X. Zeng, Y. Liu, J. Lu and X. Zhang, Nanoscale, 2020, 12, 16008–16027 RSC.
  4. H. Li, W. Wang, K. Xu, B. Cheng, J. Xu and S. Cao, Chin. J. Catal., 2025, 72, 24–47 CrossRef CAS.
  5. N. Lewis, Science, 2007, 315, 798–801 CrossRef CAS PubMed.
  6. Z. Teng, Q. Zhang, H. Yang, K. Kato, W. Yang, Y. Lu, S. Liu, C. Wang, A. Yamakata, C. Su, B. Liu and T. Ohno, Nat. Catal., 2021, 4, 374–384 CrossRef CAS.
  7. G. Duca and S. Travin, Am. J. Phys. Chem., 2020, 9, 139371 Search PubMed.
  8. J. Chang, Q. Li, J. Shi, M. Zhang, L. Zhang, S. Li, Y. Chen, S. Li and Y. Lan, Angew. Chem., Int. Ed., 2023, 62, e202218868 CrossRef CAS PubMed.
  9. K. Zhang, J. Liu, L. Wang, B. Jin, X. Yang, S. Zhang and J. Park, J. Am. Chem. Soc., 2020, 142, 8641–8648 CrossRef CAS PubMed.
  10. C. Zhao, X. Wang, S. Ye and J. Liu, Sol. RRL, 2022, 6, 2200427 CrossRef CAS.
  11. R. Service, Science, 2005, 309, 548–551 CrossRef CAS PubMed.
  12. Z. Chen, C. Chu, D. Yao, Q. Li and S. Mao, Adv. Funct. Mater., 2024, 34, 2400506 CrossRef CAS.
  13. D. Tsukamoto, A. Shiro, Y. Shiraishi, Y. Sugano, S. Ichikawa, S. Tanaka and T. Hirai, ACS Catal., 2012, 2, 599–603 CrossRef CAS.
  14. Q. Tian, X. Zeng, C. Zhao, L. Jing, X. Zhang and J. Liu, Adv. Funct. Mater., 2023, 33, 2213173 CrossRef CAS.
  15. J. Ran, W. Guo, H. Wang, B. Zhu, J. Yu and S. Qiao, Adv. Mater., 2018, 30, 1800128 CrossRef PubMed.
  16. Y. Shiraishi, T. Takii, T. Hagi, S. Mori, Y. Kofuji, Y. Kitagawa, S. Tanaka, S. Ichikawa and T. Hirai, Nat. Mater., 2019, 18, 985–993 CrossRef CAS PubMed.
  17. J. Sun, H. Jena, C. Krishnaraj, K. Rawat, S. Abednatanzi, J. Chakraborty, A. Laemont, W. Liu, H. Chen, Y. Liu, K. Leus, H. Vrielinck, V. Speybroeck and P. Voort, Angew. Chem., Int. Ed., 2023, 62, e202216719 CrossRef CAS PubMed.
  18. C. Zhang, L. Yuan, C. Liu, Z. Li, Y. Zou, X. Zhang, Y. Zhang, Z. Zhang, G. Wei and C. Yu, J. Am. Chem. Soc., 2023, 145, 7791–7799 CrossRef CAS PubMed.
  19. X. Zhang, P. Ma, C. Wang, L. Gan, X. Chen, P. Zhang, Y. Wang, H. Li, L. Wang, X. Zhou and K. Zheng, Energy Environ. Sci., 2022, 15, 830–842 RSC.
  20. Z. Yong and T. Ma, Angew. Chem., Int. Ed., 2023, 62, e202308980 CrossRef CAS PubMed.
  21. Y. Li, D. Luan and X. Lou, Adv. Mater., 2024, 36, 2412386 CrossRef CAS PubMed.
  22. J. Yang, X. Zeng, M. Tebyetekerwa, Z. Wang, C. Bie, X. Sun, I. Marriam and X. Zhang, Adv. Energy Mater., 2024, 14, 2400740 CrossRef CAS.
  23. W. Yu, C. Hu, L. Bai, N. Tian, Y. Zhang and H. Huang, Nano Energy, 2022, 104, 107906 CrossRef CAS.
  24. C. Zhao, H. Li, Y. Yin, W. Tian, X. Yan, J. He, Z. Chen, S. Ye and J. Liu, Angew. Chem., Int. Ed., 2025, 64, e202420895 CrossRef CAS PubMed.
  25. J. Wang, Y. Zhang, S. Jiang, C. Sun and S. Song, Angew. Chem., Int. Ed., 2023, 62, e202307808 CrossRef CAS PubMed.
  26. T. Tian, Z. Wang, K. Li, H. Jin, Y. Tang, Y. Sun, P. Wan and Y. Chen, Materials, 2024, 17, 1748 CrossRef CAS PubMed.
  27. Z. Wang, C. Li and K. Domen, Chem. Soc. Rev., 2019, 48, 2109–2125 RSC.
  28. D. Chen, W. Chen, Y. Wu, L. Wang, X. Wu, H. Xu and L. Chen, Angew. Chem., Int. Ed., 2023, 62, e202217479 CrossRef CAS PubMed.
  29. Y. Huang, M. Shen, H. Yan, Y. He, J. Xu, F. Zhu, X. Yang, Y. Ye and G. Ouyang, Nat. Commun., 2024, 15, 5406 CrossRef CAS PubMed.
  30. Y. Kofuji, Y. Isobe, Y. Shiraishi, H. Sakamoto, S. Tanaka, S. Ichikawa and T. Hirai, J. Am. Chem. Soc., 2016, 138, 10019–10025 CrossRef CAS PubMed.
  31. Y. Bo, C. Gao and Y. Xiong, Nanoscale, 2020, 12, 12196–12209 RSC.
  32. Y. Gao, X. Zhang, C. Ban, R. Feng, J. Hou, J. Meng, G. Yang, C. Gao, L. Xia, P. Ma, K. Wang and X. Qu, Mater. Today Energy, 2024, 40, 101483 CrossRef CAS.
  33. Y. Zheng, H. Zhou, B. Zhou, J. Mao and Y. Zhao, Catal. Sci. Technol., 2022, 12, 969–975 RSC.
  34. Y. Xu, H. Fu, L. Zhao, L. Jian, Q. Liang and X. Xiao, New J. Chem., 2021, 45, 3335–3342 RSC.
  35. S. Li, G. Dong, R. Hailili, L. Yang, Y. Li, F. Wang, Y. Zeng and C. Wang, Appl. Catal., B, 2016, 190, 26–35 CrossRef CAS.
  36. L. Shi, L. Yang, W. Zhou, Y. Liu, L. Yin, X. Hai, H. Song and J. Ye, Small, 2018, 14, 1703142 CrossRef PubMed.
  37. Z. Tian, C. Han, Y. Zhao, W. Dai, X. Lian, Y. Wang, Y. Zheng, Y. Shi, X. Pan, Z. Huang, H. Li and W. Chen, Nat. Commun., 2021, 12, 2039 CrossRef CAS PubMed.
  38. C. Wu, Z. Teng, C. Yang, F. Chen, H. Yang, L. Wang, H. Xu, B. Liu, G. Zheng and Q. Han, Adv. Mater., 2022, 34, 2110266 CrossRef CAS PubMed.
  39. L. Li, X. Lv, Y. Xue, H. Shao, G. Zheng and Q. Han, Angew. Chem., Int. Ed., 2024, 63, e202320218 CrossRef CAS PubMed.
  40. E. Zhou, F. Wang, X. Zhang, Y. Hui and Y. Wang, Angew. Chem., Int. Ed., 2024, 63, e202400999 CrossRef CAS PubMed.
  41. M. Teranishi, R. Hoshino, S. Naya and H. Tada, Angew. Chem., Int. Ed., 2016, 55, 12773–12777 CrossRef CAS PubMed.
  42. S. Bai, L. Wang, Z. Li and Y. Xiong, Adv. Sci., 2016, 4, 1600216 CrossRef PubMed.
  43. Z. Chen, H. Chen, K. Wang, J. Chen, M. Li, Y. Wang, P. Tsiakaras and S. Song, ACS Catal., 2023, 13, 6497–6508 CrossRef CAS.
  44. Y. Shao, Y. Zhang, C. Chen, S. Dou, Y. Lou, Y. Dong, Y. Zhu and C. Pan, Chin. J. Catal., 2024, 61, 205–214 CrossRef CAS.
  45. Y. Wang, W. Yu, C. Wang, F. Chen, T. Ma and H. Huang, eScience, 2024, 4, 100228 CrossRef.
  46. D. Maarisetty and S. S. Baral, J. Mater. Chem. A, 2020, 8, 18560–18604 RSC.
  47. F. N. Habarugira, D. Yao, W. Miao, C. Chu, Z. Chen and S. Mao, Chin. Chem. Lett., 2024, 35, 109886 CrossRef CAS.
  48. S. Bai, N. Zhang, C. Gao and Y. Xiong, Nano Energy, 2018, 53, 296–336 CrossRef CAS.
  49. W. Chi, B. Liu, Y. Dong, J. Zhang, X. Sun, C. Pan, H. Zhao, Y. Ling and Y. Zhu, Appl. Catal., B, 2024, 355, 124077 CrossRef CAS.
  50. W. Li, B. Han, Y. Liu, J. Xu, H. He, G. Wang, J. Li, Y. Zhai, X. Zhu and Y. Zhu, Angew. Chem., Int. Ed., 2024, 64, e202421356 CrossRef PubMed.
  51. X. Zhang, S. Cheng, C. Chen, X. Wen, J. Miao, B. Zhou, M. Long and L. Zhang, Nat. Commun., 2024, 15, 2649 CrossRef CAS PubMed.
  52. Z. Xie, X. Chen, W. Wang, X. Ke, X. Zhang, S. Wang, X. Wu, J. Yu and X. Wang, Angew. Chem., Int. Ed., 2024, 63, e202410179 CrossRef CAS PubMed.
  53. Y. Guo, Y. Dong, B. Liu, B. Ni, C. Pan, J. Zhang, H. Zhao, G. Wang and Y. Zhu, Adv. Funct. Mater., 2024, 34, 2402920 CrossRef CAS.
  54. Y. Lin, J. Zou, X. Wu, S. Tong, Q. Niu, S. He, S. Luo and C. Yang, Nano Lett., 2024, 24, 6302–6311 CrossRef CAS PubMed.
  55. K. Li, Q. Ge, Y. Liu, L. Wang, K. Gong, J. Liu, L. Xie, W. Wang, X. Ruan and L. Zhang, Energy Environ. Sci., 2023, 16, 1135–1145 RSC.
  56. L. Chen, C. Chen, Z. Yang, S. Li, C. Chu and B. Chen, Adv. Funct. Mater., 2021, 31, 2105731 CrossRef CAS.
  57. Y. Luo, B. Zhang, C. Liu, D. Xia, X. Ou, Y. Cai, Y. Zhou, J. Jiang and B. Han, Angew. Chem., Int. Ed., 2023, 62, e202305355 CrossRef CAS PubMed.
  58. N. Li, C. Wang, W. Gao, W. Li, B. Li and X. Wang, Sep. Purif. Technol., 2025, 356, 129978 CrossRef CAS.
  59. H. Yu, F. Zhang, Q. Chen, P. Zhou, W. Xing, S. Wang, G. Zhang, Y. Jiang and X. Chen, Angew. Chem., Int. Ed., 2024, 63, e202402297 CrossRef CAS PubMed.
  60. S. Li, R. Ma, C. Tu, W. Zhang, R. Li, Y. Zhao and K. Zhang, Angew. Chem., Int. Ed., 2024, 64, e202421040 CrossRef PubMed.
  61. Q. Zhu, L. Shi, Z. Li, G. Li and X. Xu, Angew. Chem., Int. Ed., 2024, 63, e202408041 CrossRef CAS PubMed.
  62. P. Dong, X. Xu, T. Wu, R. Luo, W. Kong, Z. Xu, S. Yuan, J. Zhou and J. Lei, Angew. Chem., Int. Ed., 2024, 63, e202405313 CrossRef CAS PubMed.
  63. R. Ma, L. Wang, H. Wang, Z. Liu, M. Xing, L. Zhu, X. Meng and F. Xiao, Appl. Catal., B, 2019, 244, 594–603 CrossRef CAS.
  64. D. Wen, Y. Su, J. Fang, D. Zheng, Y. Xu, S. Zhou, A. Meng, P. Han and C. Wong, Nano Energy, 2023, 117, 108917 CrossRef CAS.
  65. V. Maurino, C. Minero, G. Mariella and E. Pelizzetti, Chem. Commun., 2005, 2627–2629 RSC.
  66. X. Sun, J. Yang, X. Zeng, L. Guo, C. Bie, Z. Wang, K. Sun, A. K. Sahu, M. Tebyetekerwa, T. E. Rufford and X. Zhang, Angew. Chem., Int. Ed., 2024, 63, e202414417 CrossRef CAS PubMed.
  67. L. Wang, J. Zhang, Y. Zhang, H. Yu, Y. Qu and J. Yu, Small, 2021, 18, 2104561 CrossRef PubMed.
  68. C. Jin, H. Shen, J. Li, X. Guo, S. Rao, W. Yang, Q. Liu, Z. Sun and J. Yang, Nano Lett., 2024, 24, 14484–14492 CrossRef CAS PubMed.
  69. C. Zhu, Y. Yao, Q. Fang, S. Song, B. Chen and Y. Shen, ACS Catal., 2024, 14, 2847–2858 CrossRef CAS.
  70. X. Zhang, D. Gao, B. Zhu, B. Cheng, J. Yu and H. Yu, Nat. Commun., 2024, 15, 3212 CrossRef CAS PubMed.
  71. T. Liu, Z. Pan, K. Kato, J. J. M. Vequizo, R. Yanagi, X. Zheng, W. Yu, A. Yamakata, B. Chen, S. Hu, K. Katayama and C. Chu, Nat. Commun., 2022, 13, 7783 CrossRef CAS PubMed.
  72. Y. Li, Y. Guo, D. Luan, X. Gu and X. Lou, Angew. Chem., Int. Ed., 2023, 62, e202310847 CrossRef CAS PubMed.
  73. H. Jiang, L. Wang, X. Yu, L. Sun, J. Li, J. Yang and Q. Liu, Chem. Eng. J., 2023, 466, 143129 CrossRef CAS.
  74. L. Qin, P. She, P. Si, S. Liu, G. Dai and C. Pan, Environ. Res., 2025, 264, 120309 CrossRef CAS PubMed.
  75. J. Hu, B. Li, X. Li, T. Yang, X. Yang, J. Qu, Y. Cai, H. Yang and Z. Lin, Adv. Mater., 2024, 36, 2412070 CrossRef CAS PubMed.
  76. X. Chen, W. Zhang, L. Zhang, L. Feng, C. Zhang, J. Jiang, T. Yan and H. Wang, J. Mater. Chem. A, 2020, 8, 18816–18825 RSC.
  77. Y. Yang, Q. Guo, Q. Li, L. Guo, H. Chu, L. Liao, X. Wang, Z. Li and W. Zhou, Adv. Funct. Mater., 2024, 34, 2400612 CrossRef CAS.
  78. G. Moon, W. Kim, A. D. Bokare, N. Sung and W. Choi, Energy Environ. Sci., 2014, 7, 4023–4028 RSC.
  79. S. Zhao, X. Zhao, H. Zhang, J. Li and Y. Zhu, Nano Energy, 2017, 35, 405–414 CrossRef CAS.
  80. S. Zhao, X. Zhao, S. Ouyang and Y. Zhu, Catal. Sci. Technol., 2018, 8, 1686–1695 RSC.
  81. Y. Kofuji, S. Ohkita, Y. Shiraishi, H. Sakamoto, S. Tanaka, S. Ichikawa and T. Hirai, ACS Catal., 2016, 6, 7021–7029 CrossRef CAS.
  82. J. Zhang, L. Zheng, F. Wang, C. Chen, H. Wu, S. A. K. Leghari and M. Long, Appl. Catal., B, 2020, 269, 118770 CrossRef CAS.
  83. Y. Pan, W. Si, Y. Li, H. Tan, J. Liang and F. Hou, Chin. Chem. Lett., 2024, 35, 109877 CrossRef CAS.
  84. J. Cai, J. Huang, S. Wang, J. Iocozzia, Z. Sun, J. Sun, Y. Yang, Y. Lai and Z. Lin, Adv. Mater., 2019, 31, 1806314 CrossRef PubMed.
  85. R. Du, K. Xiao, B. Li, X. Han, C. Zhang, X. Wang, Y. Zuo, P. Guardia, J. Li, J. Chen, J. Arbiol and A. Cabot, Chem. Eng. J., 2022, 441, 135999 CrossRef CAS.
  86. J. Yang, D. Wang, H. Han and C. Li, Acc. Chem. Res., 2013, 46, 1900–1909 CrossRef CAS PubMed.
  87. H. Hirakawa, S. Shiota, Y. Shiraishi, H. Sakamoto, S. Ichikawa and T. Hirai, ACS Catal., 2016, 6, 4976–4982 CrossRef CAS.
  88. Z. Chen, D. Yao, C. Chu and S. Mao, Chem. Eng. J., 2023, 451, 138489 CrossRef CAS.
  89. T. Liu, Z. Pan, J. J. M. Vequizo, K. Kato, B. Wu, A. Yamakata, K. Katayama, B. Chen, C. Chu and K. Domen, Nat. Commun., 2022, 13, 1034 CrossRef CAS PubMed.
  90. Y. Wang, Y. Wang, J. Zhao, M. Chen, X. Huang and Y. Xu, Appl. Catal., B, 2021, 284, 119691 CrossRef CAS.
  91. Y. Liu, L. Li, H. Tan, N. Ye, Y. Gu, S. Zhao, S. Zhang, M. Luo and S. Guo, J. Am. Chem. Soc., 2023, 145, 19877–19884 CrossRef CAS PubMed.
  92. W. Cai, Y. Tanaka, X. Zhu and T. Ohno, Nano Res., 2024, 17, 7027–7038 CrossRef CAS.
  93. R. Zhang, H. Xu, Z. Huang, J. Zhang, L. Liu, Z. Ma, Z. Zhang, K. Wang, P. Liu, H. Liu and X. Zheng, Adv. Funct. Mater., 2024, 35, 2420504 CrossRef.
  94. H. Tan, P. Zhou, M. Liu, Y. Gu, W. Chen, H. Guo, J. Zhang, K. Yin, Y. Zhou, C. Shang, Q. Zhang, L. Gu, N. Zhang, J. Ma, Z. Zheng, M. Luo and S. Guo, J. Am. Chem. Soc., 2024, 146, 31950–31960 CrossRef CAS PubMed.
  95. Y. Li, Y. Guo, G. Fan, D. Luan, X. Gu and X. Lou, Angew. Chem., Int. Ed., 2024, 63, e202317572 CrossRef CAS PubMed.
  96. Y. Kondo, K. Honda, Y. Kuwahara, K. Mori, H. Kobayashi and H. Yamashita, ACS Catal., 2022, 12, 14825–14835 CrossRef CAS.
  97. S. Chen, Y. Qi, C. Li, K. Domen and F. Zhang, Joule, 2018, 2, 2260–2288 CrossRef CAS.
  98. S. Chen, C. Li, K. Domen and F. Zhang, Joule, 2023, 7, 2445–2467 CrossRef CAS.
  99. R. Abe, K. Sayama, K. Domen and H. Arakawa, Chem. Phys. Lett., 2001, 344, 339–344 CrossRef CAS.
  100. A. J. Bard, J. Photochem., 1979, 10, 59–75 CrossRef CAS.
  101. J. Cheng, W. Wang, J. Zhang, S. Wan, B. Cheng, J. Yu and S. Cao, Angew. Chem., Int. Ed., 2024, 63, e202406310 CrossRef CAS PubMed.
  102. K. Meng, J. Zhang, B. Cheng, X. Ren, Z. Xia, F. Xu, L. Zhang and J. Yu, Adv. Mater., 2024, 36, 2406460 CrossRef CAS PubMed.
  103. L. Wang, J. Sun, B. Cheng, R. He and J. Yu, J. Phys. Chem. Lett., 2023, 14, 4803–4814 CrossRef CAS PubMed.
  104. X. Sun, L. Li, S. Jin, W. Shao, H. Wang, X. Zhang and Y. Xie, eScience, 2023, 3, 100095 CrossRef.
  105. H. He, Z. Wang, J. Zhang, C. Shao, K. Dai and K. Fan, Adv. Funct. Mater., 2024, 34, 2315426 CrossRef CAS.
  106. D. Sundar, C. Liu, S. Anandan and J. Wu, Molecules, 2023, 28, 5383 CrossRef CAS PubMed.
  107. J. Xiong, X. Li, J. Huang, X. Gao, Z. Chen, J. Liu, H. Li, B. Kang, W. Yao and Y. Zhu, Appl. Catal., B, 2020, 266, 118602 CrossRef CAS.
  108. S. Zhao and X. Zhao, J. Catal., 2018, 366, 98–106 CrossRef CAS.

Footnote

These authors contributed equally to the work.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.