Next Article in Journal
Copper- and Manganese-Based Bimetallic Layered Double Hydroxides for Catalytic Reduction of Methylene Blue
Previous Article in Journal
Selective Hydrogenolysis of Tetrahydrofurfuryl Alcohol to 1,5-Pentanediol over MgAl2O4-Modified Pt/WO3/γ-Al2O3 Catalyst
Previous Article in Special Issue
Effect of NaOH Concentration on Rapidly Quenched Cu–Al Alloy-Derived Cu Catalyst for CO2 Hydrogenation to CH3OH
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalytic Production of Hydrogen Peroxide from Covalent-Organic-Framework-Based Materials: A Mini-Review

Department of Chemistry, Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Fudan University, Shanghai 200433, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(7), 429; https://doi.org/10.3390/catal14070429
Submission received: 3 June 2024 / Revised: 28 June 2024 / Accepted: 2 July 2024 / Published: 5 July 2024
(This article belongs to the Special Issue Exclusive Papers in Green Photocatalysis from China)

Abstract

:
Hydrogen peroxide (H2O2) is one of the most environmentally friendly and versatile chemical oxidizing agents, with only O2 and H2O as reaction products. It is widely used in environmental protection, industrial production, and medical fields. At present, most of the industrial production of H2O2 adopts anthraquinone oxidation, but there are shortcomings such as pollution of the environment and large energy consumption. Covalent organic frameworks (COFs) are a class of porous crystalline materials formed by organic molecular building blocks connected by covalent bonds. The ordered conjugated structure of COFs not only facilitates the absorption of light energy but also promotes the transport of excited-state electrons. Therefore, the photochemical synthesis of H2O2 from water and oxygen using photocatalysts based on COFs as a green route has attracted much attention. In this review, we provide an overview of recent studies on COFs as photocatalysts and the different mechanisms involved in the photocatalytic production of hydrogen peroxide. Then, we summarize the various strategies to improve the performance. Finally, we outline the challenges and future directions of COFs in practical applications. This review highlights the potential and application prospects of COFs in the photochemical synthesis of H2O2, aiming to provide guidance for the design of COF-based catalysts and the optimization for photocatalytic production of H2O2, in order to promote scientific development and application in this field.

Graphical Abstract

1. Introduction

As a clean and inexhaustible source of energy, solar energy is a good substitute for traditional fossil energy. Photocatalytic technology is a significant milestone in the exploration of sustainable energy, providing new ideas for stimulating chemical reactions [1,2]. Photocatalytic synthesis of hydrogen peroxide is a significant example that could provide a green and safe route to the field of H2O2 synthesis. H2O2 was first synthesized by Louis Jacques Thénard in 1818 [3]. Since hydrogen peroxide is regarded as one of the cleanest and universal chemical oxidants [4], with only O2 and H2O as reaction products, the scope of its application is constantly expanding, being widely used in environmental protection, industrial production, and medicine [5,6,7].
Currently, the dominant method for industrial production of hydrogen peroxide is the anthraquinone oxidation (AO) process, which includes the hydrogenation of anthraquinone followed by oxidation of the obtained dihydroanthraquinone to produce H2O2. However, the AO process is associated with high costs due to the extraction solvent recovery, hydrogenation catalyst, and expensive quinone required [8,9,10]. In addition, this method has limitations such as unsuitability for daily on-site use, environmental pollution, and significant energy consumption, which impede the wide application of hydrogen peroxide as a green oxidant and fuel. In an alternative method, H2O2 can be directly synthesized from mixtures of H2 and O2 [11,12]. However, one major challenge lies in the precise control of the hydrogen-to-oxygen ratio. The direct oxidation reaction of H2O and O2 at high temperatures and pressures increases the risk of safety hazards such as explosions and fires. As a result, a safe and environmentally friendly method for producing H2O2 is urgently needed.
In recent years, the process of photocatalytic production of H2O2 has received increasing attention [13,14], providing a new perspective on H2O2 synthesis. Previous studies have synthesized various photocatalysts for photocatalytic hydrogen peroxide production. Currently, typical catalyst materials include graphitic carbon nitride (g-C3N4) [15], metal oxides (TiO2, WO3, ZnO) [16,17], metal sulfides (ZnS, CdS, In2S3) [18,19,20], and organic frameworks (MOF and COF) [21,22,23]. In 2014, Hirai and co-workers [24] demonstrated the feasibility of producing H2O2 with the metal-free polymer photocatalyst g-C3N4 for the first time. Despite the good selectivity and innovativeness of g-C3N4, its fast electron–hole complexation and insufficient light absorption result in a low H2O2 yield (only 30 μmol of H2O2 is formed in 12 h of reaction) with ethanol as a sacrificial agent. Since then, researchers have been exploring different methods to exploit the potential of g-C3N4 photocatalytic hydrogen peroxide production. Compared to inorganic photocatalysts, organic photocatalysts offer a wider range of precise chemical and electronic modifications at the molecular level due to the diverse constituent units and superior adjustability. COFs were first discovered by Yaghi and colleagues in 2005 [25]. As novel metal-free polymer porous materials, COFs have attracted significant attention owing to their high specific surface area, uniform porosity, and adjustable structural properties [26]. Over the past two decades, researchers have focused on constructing functional COFs with desired properties and expanding their applications.
In 2020, Van Der Voort and co-authors successfully used COFs as photocatalysts to produce H2O2 [22], opening up new possibilities for metal-free photocatalysis powered by solar energy. Different from inorganic catalysts, metal-free polymer photocatalysts usually exhibit better H2O2 generation performance, probably because of their high selectivity for two-electron O2 reduction and low activity for H2O2 decomposition [27]. Since then, there has been a boom in the study of COFs for photocatalytic hydrogen peroxide production (Table 1). Through in-depth study of structure and properties, it is expected that highly efficient and stable COF-based photocatalysts will be developed to provide a new solution for photocatalytic hydrogen peroxide generation.
However, in recent years, it has been challenging for researchers to screen and summarize the information from the large number of research reports in the field of photocatalytic H2O2 production from COFs. In this review, we intend to elaborate on the research results of COFs in H2O2 photocatalytic synthesis in the last three years from the perspectives of the reaction mechanism, COF design, and optimization. Following this, the main challenges in the current research as well as the future research directions are analyzed, aiming to advance the further development of COF-based photocatalysts for practical H2O2 synthesis applications.

2. Principles of Photocatalytic COFs for H2O2 Production

The production of H2O2 from solar energy is a complex multistep process involving multiple aspects of photophysics and photochemistry. The process is divided into three stages: photon absorption, carrier dynamics, and surface reactions (Figure 1). Initially, after the photocatalyst absorbs photons with higher energy than the band gap, electrons (e) are excited into the conduction band (CB) and holes (h+) remain in the valence band (VB), forming electron-hole pairs. Subsequently, these electron-hole pairs separate into free carriers, which partly undergo a combination of the bulk phase and partly migrate to the surface of the photocatalyst. Finally, if the free carriers migrating to the surface can overcome the overpotential of each electron transfer step, they can drive the oxygen reduction (ORR) or water oxidation (WOR) half-reaction to produce H2O2, with some of the surface carriers undergoing combination. The redox capacity of electrons and holes is closely related to the conduction band minimum (CBM) and valence band maximum (VBM) [28,42,43]. Currently, the field of photocatalytic hydrogen peroxide production is mainly limited by poor solar energy response and high photogenerated carrier combination efficiency. COFs, due to their large surface area, large building units in the skeleton, and abundant bonding reactions, provide abundant catalytic sites and a large space for precise control of their electronic structures by adjusting their conformations to satisfy the needs of various redox reactions [44,45]. In this section, we present COF-based photocatalysts for the synthesis of H2O2 via different reaction pathways.
Theoretically, we can classify COF-based catalysts into three categories by the pathway of photocatalytic hydrogen peroxide production: catalysts specifically targeting ORR catalysts, specifically targeting WOR catalysts, and catalysts through dual-channel pathway [46] (Figure 1). However, due to the high structural stability of the H2O molecule, a large activation energy barrier needs to be overcome to achieve the 2e water oxidation process for H2O2 production. This results in a relative scarcity of COF catalysts specifically designed to produce H2O2 via the 2e WOR process in practice. For practical applications in this research area, COF photocatalysts for photocatalytic generation of H2O2 can be mainly classified into two main types: catalysts generating H2O2 mainly via the ORR and those via dual-channel pathway with ORR and WOR. In this section, we introduce the latest research on these two types of COF catalysts via different reaction paths. Different functional group modifications and different active sites result in different photocatalytic pathways. For example, in the case of ORR, commonly used active sites include benzene, pyrene, diarylamine, pyridine, and so on. For WOR, active sites include triazine, acetylene, diacetylene, and so on.

2.1. Photocatalytic COFs Mainly via ORR Pathway

The oxygen reduction reaction is easier to achieve in photocatalysis than the water oxidation reaction. The current photocatalytic generation of hydrogen peroxide using COF-based photocatalysts mainly depends on the ORR process. The ORR process leads to the generation of H2O2 by two routes: the direct ORR route involving two electron steps and the two-step 1e- ORR route via a proton-coupled electron transfer (PCET) process (Figure 1). In the PCET process, protons are mainly derived from water molecules or organic electron donors. Compared to the consecutive two-step 1e ORR pathway, the direct one-step 2e ORR pathway exhibits more beneficial thermodynamic energy levels (O2/H2O2 0.68 VNHE vs. O2/O2 −0.33 VNHE) [29,47].
Moreover, the direct one-step 2e ORR process avoids the formation of highly reactive O2 intermediates, which reduces side reactions. The O2 intermediates are susceptible to aqueous matrixes, such as Cl and dissolved organic matter in natural water, thus decreasing the efficiency of H2O2 production in actual applications [48]. Therefore, regulating the ORR process in the one-step 2e pathway is important for efficient and highly selective in situ H2O2 synthesis. Xi and co-authors [30] synthesized a series of COFs modified with different diazines (include pyrazine, pyrimidine, and pyridazine) for H2O2 generation without the use of sacrificial agents (Figure 2a). As pyridazine is embedded in TPDz, it stabilizes endoperoxide intermediate species, improving the efficiency of the direct 2e ORR process (Figure 2b,c). Compared to pyridazine, the stepwise 1e pathway performs a more essential function in the photosynthesis of H2O2 in TpMd- and TpPz-based photocatalytic processes. The results showed that TpDz via the direct 2e ORR process had the highest yield of 7327 μmol h−1g−1 (Figure 2d), which was one of the best photocatalysts for H2O2 production from COF-based catalysts, indicating the role of the reaction pathway in the catalytic performance.

2.2. Photocatalytic COFs via Dual-Channel Pathway

Photocatalytic H2O2 synthesis via the ORR and WOR dual-channel pathways is a non-polluting and sustainable method with greater atom utilization compared with the process mainly via ORR [31]. To achieve highly efficient photosynthesis of H2O2, WOR is essential to balance the whole photocatalytic system. However, due to the high thermodynamic requirements of WOR, this process is relatively difficult to realize. Previous studies have often relied on the use of sacrificial agents, such as methanol, to reduce the energy barrier of WOR [32,49]. Nevertheless, the use of sacrificial agents may introduce impurities and generate additional by-products, which limits the practical application of photocatalysts and the purity of H2O2. Therefore, the development of COFs capable of efficiently generating H2O2 via a dual-channel pathway without sacrificial agents is of great scientific significance and application value. Tang and co-workers [33] successfully constructed two new thiophene-containing COFs, TD-COF and TT-COF (Figure 3a), and used them as catalysts for photocatalytic production of H2O2. Highly efficient synthesis of H2O2 was achieved under visible light irradiation through the two-step 1e- ORR and direct 2e WOR pathways. The performance of photocatalytic H2O2 generation was optimized by adapting the n-heterocyclic units such as pyridine and triazine in the COFs. The H2O2 yields of TD-COF reached 4060 μmol h−1g−1 and 3364 μmol h−1g−1 in deionized water and natural seawater, respectively, without any sacrificial agents (Figure 3b–d). DFT calculations revealed the thiophene unit as the main active center for photoreduction, whereas the benzene ring, which is connected through imine bonding, is the active center for photo-oxidation.

3. Strategies to Improve Photocatalytic H2O2 Production

In this section, we present strategies to improve photocatalytic H2O2 production. COF-based photocatalysts have a comparatively broad response spectrum, easily adjustable electronic structure, high specific surface area, and diverse pore structure in the field of H2O2 production, which are promising for their applications. Most COFs are competent for generating hydrogen peroxide without a sacrificial agent due to the presence of a conjugated structure that favors electron transfer. However, the concentration of H2O2 produced at the current level of research is still comparatively low, which makes COFs unsuitable for practical use. The efficiency of H2O2 generation can be significantly improved by strategies such as functional group modification, design of redox center-separated COFs, and synthesis of COF nanohybrids.

3.1. Functional Groups Modification

The incorporation of branched functional groups into organic semiconductors is generally regarded as an excellent method of imparting unique properties to these semiconductors [50,51]. Functionalization is an important means of adapting and optimizing the properties of COFs involving pre-synthetic or post-synthetic modifications to the organic backbone or pore interior of the COFs. The introduction of functional groups (hydroxy group [52], amine group [53], and acetylene [54]) into COFs can change the inner electron arrangement and effectively solve the problem of rapid combination of the inner electrons and holes, thus enhancing the photocatalytic hydrogen peroxide production activity [55]. Han and co-authors designed an excellent photocatalyst by modifying sulfone units into COFs (FS-COFs) (Figure 4a) [29]. Sulfone units in FS-COFs accelerated the separation of photogenerated carriers, facilitating the protonation and promoting the adsorption of oxygen (Figure 4b). The FS-COFs achieved highly selective and efficient H2O2 generation through the 2e ORR process (Figure 4c) with a hydrogen peroxide yield of 3904.2 μmol h−1g−1 under visible light irradiation.
The introduction of functional groups not only promotes the separation of light-induced electron-hole pairs but also contributes to the overall catalytic system. For instance, Guo et al. reported a new fluorinated covalent organic framework strategy to enhance the limitation of palladium metal-isolated clusters (Pd-ICs) by COFs for enhanced photosynthesis of H2O2 [34]. The introduction of strongly electronegative fluorine into TAPT-TFPA COFs enhances the interaction between COFs and Pd-ICs, which can optimize the d-band center, thus improving the performance of photocatalytic synthesis of H2O2 (Figure 4d). The prepared TAPT-TFPA COFs@Pd-ICs have an efficient photocatalytic H2O2 yield that can reach 2143 μmol h−1g−1. More significantly, the stability of TAPT-TFPA COFs@Pd-ICs in the photocatalytic H2O2 production process is greatly improved, and the photocatalytic activity can remain stable after more than 100 h (Figure 4e), which is the best stability among reported COF-based photocatalytic materials.

3.2. Design of Spatially Separated Redox Centers

According to the above discussion, photocatalytic hydrogen peroxide production can be achieved by oxygen reduction half-reaction (ORR) and water oxidation half-reaction (WOR). If the ORR and WOR occur at the same or close catalytic sites, the photogenerated electrons and holes involved in the half-reactions will be easily recombined, and the overall catalytic performance will be limited. Hence, spatial separation of the oxidation and reduction centers to inhibit charge recombination is an excellent method to improve the efficiency of H2O2 generation [56,57]. This design strategy takes advantage of the structural customizability and functional tunability of COFs. By introducing different organic building blocks into the skeleton of COFs, the redox centers are spatially separated so that the electrons and holes are confined to different regions, which facilitates the effective separation and transfer of charge carriers (Figure 5a). Such a design can increase the lifetime of light-generated electrons and holes, thus enhancing the efficiency of the photocatalytic reaction [58]. Chen and co-authors [35] reported s-heptazine-based COFs (Figure 5b), which possessed separated redox centers and were able to efficiently catalyze the production of hydrogen peroxide from O2 and pure water via a dual-channel pathway without the addition of sacrificial agents. The spatially separated redox centers of HEP-TAPB-COF and HEP-TAPT-COF efficiently promote charge separation, which improves the performance of photocatalytic production of H2O2. The triazine and s-heptazine groups in HEP-TAPT-COF constitute a double active center to produce hydrogen peroxide (87.50 μmol  h−1), which has better performance than HEP-TAPB-COF with only the s-heptazine unit involved in the generation of H2O2 (49.50 μmol  h−1).
Wang et al. [36] designed benzotrithiophene (Btt)-based COFs with spatially separated redox centers for the photocatalytic synthesis of H2O2 from H2O and O2 without sacrificial agents (Figure 6a,b). DFT calculations showed that the binding of Btt to various functional groups (Tpa, Tapb, and Tapt) modulates the electron distribution on the C atom near the imine bond, thus enhancing the two-step 1e- oxygen reduction reaction by facilitating the in-plane charge transfer and the binding intensity of the O2* and OOH* intermediates. This structural design achieves spatial separation of the redox centers in the COF, thus facilitating efficient charge separation and transfer. The yield of TaptBtt to H2O2 was 1407 μmol h−1g−1 with a photochemical conversion efficiency of 0.296%.

3.3. COF Nanohybrids

As most covalent organic frameworks are composed entirely of non-metallic elements, the centers of catalytic activity are limited, resulting in low catalytic efficiency. Numerous studies have demonstrated that combining COFs with different types of nanomaterials such as metals (Au [59]), metal sulfides (In2S3 [60], CdS [61]), metal oxides (TiO2 [62], ZnO [63], WO3 [64]), metal organic frameworks (MOFs [63]), and carbon-based materials (g-C3N4 [65], CQDs [66]) can significantly enhance the efficiency of COFs in photocatalytic reactions. This combination not only greatly enhances the light absorption of COF-based photocatalysts but also forms a heterojunction structure for enhanced photocatalytic performance [67]. Therefore, the development of COF-based heterostructures is not only scientifically prospective but also shows extensive potential and value in practical photocatalytic hydrogen peroxide production reactions. Lu and co-authors synthesized floral inorganic/organic (CdS/TpBpy) composites as an s-type heterojunction using a simple solvothermal–hydrothermal method [37] (Figure 7a–d). The unique floral structure increases the active sites and O2 absorption rate, and the s-type heterojunction facilitates the photogenerated carrier transfer in the internal embedded electric field. The optimal CdS/TpBpy H2O2 yield (3600 μmol  h−1) is 2.4 and 25.6 times higher than that of TpBpy and CdS, respectively, without sacrificial reagents (Figure 7e–g). What is more, for CdS/TpBpy the decomposition of H2O2 was suppressed, thus raising the total yield. Wang and co-authors synthesized S-scheme BiOBr/COF photocatalysts by in situ growth of BiOBr nanosheets (NSs) on COFs with large π-conjugated structures [68]. The highest performance of H2O2 production by BiOBr/COF photocatalysis was 3749 μmol h−1g−1, which was about 2 and 27 times higher than the H2O2 production rates of COF and BiOBr, respectively. The construction of s-type heterojunctions contributes to effective photogenerated carrier separation and transfer, thereby improving redox efficiency. In addition, the lying-down O2 adsorption configuration on the BiOBr/COF surface facilitates 2e ORR, which extremely reduces the reduction potential demand for the conversion of oxygen to hydrogen peroxide and favors the reaction. In this system, the synergistic effect between the s-type heterojunction and the apparent interaction between O2 and COF enhance the activity of photocatalytic generation of H2O2.

3.4. Linkage Chemistry

COFs are organic monomers linked together by strong covalent bonds, and more than two dozen different linkages have been developed to construct COFs. Among them, triazine, C=C, and imine are common connecting bonds. At present, most COFs are synthesized by solvothermal methods under vacuum or inert atmosphere conditions [69]. Furthermore, microwave synthesis and mechanical synthesis can also synthesize COFs, which are characterized by short reaction times and high yields [70]. Moreover, multi-component reaction (MCR) [64] synthesis methods can promote structural diversification of the constructs. In past studies, it has been shown that linkage chemistry determines COF electronic properties and catalytic behavior. Tang et al. designed a pair of isomeric COFs (TB-COF and TA-COF) with anti-substituted amine bonds [71]. After a series of performance evaluations, it was found that TB-COF exhibits significant advantages over TA-COF in H2O2 photosynthesis. This is mainly attributed to the larger dipole moment and superior carrier separation efficiency of TB-COF. Further theoretical calculations reveal the internal structural properties of TB-COF, in which the hydroxyl-rich benzene ring assumes the role of photo-oxidation, while the triazine unit is responsible for the photoreduction process. This finding reveals the possibility of optimizing the electronic structure and energy barriers of COFs by adjusting the bond orientation to effectively enhance the photocatalytic activity, which provides a new strategy and guidance for the design of efficient photocatalytic COF materials. Yu et al. constructed o-COFs and p-COFs with ortho- and para-linkages [72]. O-COFs have higher hydrogen peroxide generation efficiency and stability compared to p-COFs. Furthermore, o-COFs with ortho-linkages are more robust than those with para-linkages, due to the lower energy barrier of o-COFs in the 4e WOR that determines the rate step. This finding highlights the critical role of the linkage position in COFs in facilitating H2O2 generation.

3.5. Dimensions Control

Tuning the dimensions of the COF can make the COF-based material exhibit different properties. Although the topological diversity of COFs has been extensively studied, most COFs consist of two-dimensional (2D) networks stacked on top of each other. For 2D COFs, which are usually stacked in a face-to-face manner, some of the structures in neighboring layers have strong Π–Π interactions. Compared to 2D COFs with laminated structures, 3D COFs usually outperform 2D COFs and show superior competitiveness in terms of performance by virtue of their significantly larger specific surface area, tightly interconnected channel structure, full exposure of the functional part, and highly flexible tunability. In contrast, the dense interlayer stacking of 2D COFs results in a large number of potential active sites being “buried” by interlayer interactions, limiting their efficiency in photocatalytic hydrogen peroxide production [73,74]. However, the structural diversity of 3D COFs is greatly limited by the lack of multifaceted three-dimensional molecular building blocks and crystallization difficulties [75]. Wang et al. demonstrated a study on the synthesis and property characterization of 2D-PdPor-COF and 3D-PdPor-COF to consider the effect of dimensionality on the COF properties [76]. The results show that 3D-PdPor-COF has superior photocatalytic hydrogen peroxide production activity compared with 2D-PdPor-COF. In addition, the pore size of 3D-PdPor-COF is considerably smaller than those of 2D-PdPor-COF, and 3D-PdPor-COF has better size selectivity in photocatalytic reactions. Therefore, three-dimensional COFs are expected to be more promising photocatalysts for the photocatalytic synthesis of H2O2.
For example, Zhang et al. designed octa-aldehyde difluorobenzene monomer as 8-connected cubic nodes using a bottom-up [8 + 4] mesh strategy for the construction of three-dimensional COFs with a Tty topology (COF-NUST-16) for the first time (Figure 8a,b) [38]. An 8-connected 3D COF exhibits high photocatalytic activity because of its wide absorption range and high absorption capacity, and it is a superior dual-channel photocatalyst with a hydrogen peroxide production rate of 1081 μmol h−1g−1 compared to 2D COFs (234 μmol h−1g−1). The hydrogen peroxide production activity of 3D COFs is four times higher than 2D COFs with similar chemical structures, which demonstrates the superiority of the 3D COFs.
However, despite the significant theoretical advantages of 3D COFs, they have been reported relatively rarely in practice, mainly due to the challenge of precise structural control during synthesis. Precise control of pore size and the three-dimensional network structure of inter-pore connections requires highly specialized and sophisticated synthesis strategies and techniques, which place higher demands on experimental design and operation.

4. Perspective and Outlook

In summary, significant advancements have been accomplished in the field of photocatalytic hydrogen peroxide production using covalent organic frameworks (COFs) in the last several years. COFs, with high crystallinity and designed structures, are ideal candidates for efficient photocatalytic systems. Despite the positive outlook, the field is still in its initial stages and is still far from commercial application. There is an urgent need to address a number of issues to advance scientific research and practical applications. After reviewing current technological developments, we offer the following suggestions.

4.1. Clarify the Reaction Mechanism and the Effect of Structure on Catalytic Performance

This includes determining the pathways of hydrogen peroxide production by photocatalysis of different structures, photogenerated carrier dynamics, hydrophobicity/hydrophilicity of COFs, and adsorption of oxygen by COFs. Various studies have demonstrated that different functional groups in COFs are used as active sites for different reaction pathways [77]. For instance, bipyridine is mainly used as an ORR site [78], and acetylene is mainly used as a WOR site [31]. More research is needed in the future to determine the role of multiple functional groups in the photocatalytic production of hydrogen peroxide. Analyzing reaction mechanisms relies mainly on a number of characterization techniques. While indirect characterization techniques such as X-ray photoelectron spectroscopy (XPS), electron spin resonance (ESR) techniques, and photoluminescence (PL) spectroscopy provide useful proof for the kinetics of the photogenerated carriers and the presence of reactive intermediates, they do not provide direct evidence of the real-time involvement of electrons/holes (e/h+) and the reactive intermediates in the redox reactions at the surface. Therefore, further development of in situ characterization techniques is required to accurately provide proof of reaction mechanisms. For example, an in situ environmental transmission electron microscope (ETEM) [79] provides a way to observe the structural changes in photocatalysts at the microscopic level. In situ Fourier-transform infrared spectroscopy (FT-IR) allows for real-time monitoring of the formation and adsorption/desorption processes of intermediates on the catalyst surface. The integration of these in situ techniques will help to uncover the underlying reaction mechanisms and provide guidance for the appropriate design of COFs for efficient H2O2 production.

4.2. Improve Catalytic Performance and Applicability

This includes photocatalytic activity, stability, selectivity, and applicability in different water bodies. Enhancing the photocatalytic activity of the catalysts was mainly achieved by enhancing light absorption, promoting carrier separation and inhibiting carrier recombination. Various strategies have been developed, such as the design of redox center-separated COFs, to inhibit carrier recombination and the growth of other semiconductor photocatalyst materials on COFs to enhance light absorption and promote carrier separation. However, the activity is still far from practical application, and better strategies are needed to improve the overall activity in the future. A good photocatalyst must have high stability. Compared with inorganic photocatalysts, COFs linked by covalent bonds obviously have higher structural stability. However, the structure of COFs can be easily damaged by H2O2 with oxidizing and reducing properties as the concentration of H2O2 in the reaction system increases during the photocatalytic process, which is the major reason why a number of COF materials exhibit unstable performance in the photocatalytic synthesis of H2O2. As a result, there is a need to further explore additional improvement methods, such as functionalized modifications, formation of heterojunctions, and establishment of multiphase reaction systems to enhance the stability of COFs during the reaction process. In addition, the synthesis method also has a higher influence on the catalytic performance. Different synthesis strategies can significantly change the crystallinity, pore structure, and specific surface area of COFs which are directly related to the distribution of active sites and substrate contact efficiency during catalysis. For example, solvothermal synthesis provides more active sites and improves catalytic activity and stability, while microwave-assisted synthesis helps to obtain materials with large specific surface area and high porosity. Moreover, the development of simple synthesis methods is very important. Such methods can reduce the difficulty of synthesizing COFs and accelerate the preparation process, such as the simple synthesis method of COFs based on the imine exchange reaction [80] and the simple solution-refluxing synthesis method of 3D COF-300 [81], which contributes to the industrial application of COFs. Furthermore, the application range of COF photocatalytic hydrogen peroxide production should be improved. For instance, the pure water system is replaced by a simulated or real seawater system [82]. Seawater is a widely available resource, and the use of seawater as a reaction medium reduces dependence on freshwater resources, which contributes to sustainable development. Furthermore, it has recently been shown that different synthesis methods lead to different catalytic properties [83].

4.3. Standardize Performance Judgement Criteria

The standardization of COF photocatalytic hydrogen peroxide production performance evaluation criteria is a key step in promoting the application of COF-based materials in photocatalysis. When evaluating the photocatalytic performance of COF-based materials, there does not yet exist a universally accepted standard protocol or benchmark to ensure consistency between different studies and comparability of results. However, there are some key evaluation criteria, including H2O2 generation rate, H2O2 concentration, H2O2 decomposition, apparent quantum yield (AQY), and solar-to-chemical energy conversion efficiency (SCC efficiency). These criteria provide a quantitative approach to photocatalytic H2O2 generation performance. The judgment criteria can be regulated by standardizing the test conditions and determining the key indicators for evaluating the photocatalytic performance of COFs. What is more, there are no standardized judgement criteria for detecting the yield of H2O2 due to the difficulty of H2O2 separation and instability. Currently, the main methods for the detection of H2O2 are UV spectrophotometry and titration. However, for the same reaction system, the results obtained by different testing approaches may be very different. Authoritative detection methods and standards are needed for photocatalytic production of H2O2 driven by COF-based materials. Considering the actual production process and economic benefits, it is essential to choose the appropriate reaction conditions. Firstly, the light source is recommended to be simulated sunlight or LED lamps, which not only ensures the green sustainability of the reaction but also reduces energy consumption by effectively using natural or artificial light sources. Meanwhile, air and seawater are selected as the reaction medium, which can significantly reduce the dependence on pure oxygen and freshwater resources. In terms of temperature control, conducting photocatalytic reactions at room temperature can avoid the need for additional heating processes, thereby reducing energy consumption and production costs. In addition, the design of a simple catalyst recovery and recycling process not only reduces the cost of catalyst usage but also reduces the environmental impact of the production process. Through these combined measures, the COF photocatalytic H2O2 production process can be optimized to achieve the goal of cost-effective and environmentally friendly production.

4.4. Combine with Other Fields

In the research in the field of photocatalytic H2O2 production by COFs, the photocatalytic performance of COFs can be significantly enhanced by combining multidisciplinary knowledge from materials science, computational chemistry, and machine learning. Firstly, using innovations in materials science, novel COFs with specific functionalities can be designed and synthesized to regulate their electronic structure and pore properties through precise control at the molecular level in order to optimize light absorption and carrier transport efficiency. Secondly, applications of computational chemistry can simulate and predict the photocatalytic mechanisms of COFs, providing a theoretical basis for experimental design, such as predicting the electronic properties and reaction pathways of materials using first-principles calculations. Considering the increasing application of artificial intelligence and machine learning in materials science, we can foresee that these technologies will transform the design and synthesis of COFs. Through machine learning algorithms, researchers are able to analyze large amounts of existing data, identify key parameters that affect the performance of COFs, and predict the potential performance of new structures. For example, the crystallinity of COFs is optimized through artificial intelligence. Although this review has explored a variety of strategies to improve the photocatalytic performance of COFs, the potential for the application of interdisciplinary approaches remains enormous.

4.5. Assess the Environmental and Safety Performance of COFs

Due to their high specific surface area, high porosity, and designable chemical composition, COFs are able to efficiently utilize light energy for catalytic reactions, reducing energy consumption and the use of chemical additives and thus potentially lowering the burden on the environment. The chemical and thermal stability of COFs contributes to the safety of the reactions and reduces the generation of harmful by-products. Although there is a lack of extensive research directly addressing the life cycle assessment (LCA) of COFs, it is expected that their lower environmental footprint and potential recyclability or degradability will further enhance their environmental friendliness. However, further life cycle assessment and risk analysis studies are needed to fully assess the environmental and safety performance of COFs.

Author Contributions

J.M. and Y.H. conducted the information search and wrote the first draft of the manuscript, and X.W., Y.L., H.Z. and W.D. discussed and revised parts of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China (2021YFA1501404), the Natural Science Foundation of Shanghai (22ZR1404200), and the Natural Science Foundation of Shanghai Science and Technology Committee (19DZ2270100).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, Q.; Gu, H.; Wang, X.; Li, L.; Zhang, J.; Zhang, H.; Li, Y.-F.; Dai, W.-L. Robust hollow tubular ZnIn2S4 modified with embedded metal-organic-framework-layers: Extraordinarily high photocatalytic hydrogen evolution activity under simulated and real sunlight irradiation. Appl. Catal. B Environ. 2021, 298, 120632. [Google Scholar] [CrossRef]
  2. Zhang, H.; Gu, H.; Huang, Y.; Wang, X.; Gao, L.; Li, Q.; Li, Y.; Zhang, Y.; Cui, Y.; Gao, R.; et al. Rational design of covalent organic frameworks/NaTaO3 S-scheme heterostructure for enhanced photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2024, 664, 916–927. [Google Scholar] [CrossRef]
  3. Jose, M.C.-M.; Gema, B.-B.; Jose, L.G.F. Hydrogen peroxide synthesis: An outlook beyond the anthraquinone process. Angew. Chem. Int. Ed. 2006, 45, 6962–6984. [Google Scholar] [CrossRef]
  4. Wu, Q.; Liu, Y.; Cao, J.; Sun, Y.; Liao, F.; Liu, Y.; Huang, H.; Shao, M.; Kang, Z. A function-switchable metal-free photocatalyst for the efficient and selective production of hydrogen and hydrogen peroxide. J. Mater. Chem. A 2020, 8, 11773–11780. [Google Scholar] [CrossRef]
  5. Teong, S.P.; Li, X.; Zhang, Y. Hydrogen peroxide as an oxidant in biomass-to-chemical processes of industrial interest. Green Chem. 2019, 21, 5753–5780. [Google Scholar] [CrossRef]
  6. Fukuzumi, S.; Lee, Y.-M.; Nam, W. Recent progress in production and usage of hydrogen peroxide. Chin. J. Catal. 2021, 42, 1241–1252. [Google Scholar] [CrossRef]
  7. Chen, Z.; Yao, D.; Chu, C.; Mao, S. Photocatalytic H2O2 production Systems: Design strategies and environmental applications. Chem. Eng. J. 2023, 451, 138489. [Google Scholar] [CrossRef]
  8. García-Serna, J.; Moreno, T.; Biasi, P.; Cocero, M.J.; Mikkola, J.-P.; Salmi, T.O. Engineering in direct synthesis of hydrogen peroxide: Targets, reactors and guidelines for operational conditions. Green Chem. 2014, 16, 2320–2343. [Google Scholar] [CrossRef]
  9. Perry, S.C.; Pangotra, D.; Vieira, L.; Csepei, L.-I.; Sieber, V.; Wang, L.; Ponce de León, C.; Walsh, F.C. Electrochemical synthesis of hydrogen peroxide from water and oxygen. Nat. Rev. Chem. 2019, 3, 442–458. [Google Scholar] [CrossRef]
  10. Gao, G.; Tian, Y.; Gong, X.; Pan, Z.; Yang, K.; Zong, B. Advances in the production technology of hydrogen peroxide. Chin. J. Catal. 2020, 41, 1039–1047. [Google Scholar] [CrossRef]
  11. Edwards, J.K.; Solsona, B.N.E.N.; Carley, A.F.; Herzing, A.A.; Kiely, C.J.; Hutchings, G.J. Switching off Hydrogen Peroxide Hydrogenation in the Direct Synthesis Process. Science 2009, 323, 1037–1041. [Google Scholar] [CrossRef] [PubMed]
  12. Yang, S.; Verdaguer-Casadevall, A.; Arnarson, L.; Silvioli, L.; Čolić, V.; Frydendal, R.; Rossmeisl, J.; Chorkendorff, I.; Stephens, I.E.L. Toward the Decentralized Electrochemical Production of H2O2: A Focus on the Catalysis. ACS Catal. 2018, 8, 4064–4081. [Google Scholar] [CrossRef]
  13. Hou, H.; Zeng, X.; Zhang, X. Production of hydrogen peroxide by photocatalytic processes. Angew. Chem. Int. Ed. 2020, 59, 17356–17376. [Google Scholar] [CrossRef] [PubMed]
  14. Sun, Y.; Han, L.; Strasser, P. A comparative perspective of electrochemical and photochemical approaches for catalytic H2O2 production. Chem. Soc. Rev. 2020, 49, 6605. [Google Scholar] [CrossRef] [PubMed]
  15. Chen, J.; Kang, N.; Fan, J.; Lu, C.; Lv, K. Carbon nitride for photocatalytic water splitting to produce hydrogen and hydrogen peroxide. Mater. Today Chem. 2022, 26, 101028. [Google Scholar] [CrossRef]
  16. Kim, K.; Park, J.; Kim, H.; Jung, G.Y.; Kim, M.-G. Solid-Phase Photocatalysts: Physical Vapor Deposition of Au Nanoislands on Porous TiO2 Films for Millimolar H2O2 Production within a Few Minutes. ACS Catal. 2019, 9, 9206–9211. [Google Scholar] [CrossRef]
  17. Jiang, Z.; Cheng, B.; Zhang, Y.; Wageh, S.; Al-Ghamdi, A.A.; Yu, J.; Wang, L. S-scheme ZnO/WO3 heterojunction photocatalyst for efficient H2O2 production. J. Mater. Sci. Technol. 2022, 124, 193–201. [Google Scholar] [CrossRef]
  18. Zhang, Y.; Huang, H.; Wang, L.; Zhang, X.; Zhu, Z.; Wang, J.; Yu, W.; Zhang, Y. Cooperation of congenital and acquisitus sulfur vacancy for excellent photocatalytic hydrogen peroxide evolution of CdS nanorods from air. Chem. Eng. J. 2022, 454, 140420. [Google Scholar] [CrossRef]
  19. Li, F.; Tang, X.; Hu, Z.; Li, X.; Li, F.; Xie, Y.; Jiang, Y.; Yu, C. Boosting the hydrogen peroxide production over In2S3 crystals under visible light illumination by gallium ions doping and sulfur vacancies modulation. Chin. J. Catal. 2023, 55, 253–264. [Google Scholar] [CrossRef]
  20. Liu, C.; Bao, T.; Yuan, L.; Zhang, C.; Wang, J.; Wan, J.; Yu, C. Semiconducting MOF@ZnS Heterostructures for Photocatalytic Hydrogen Peroxide Production: Heterojunction Coverage Matters. Adv. Funct. Mater. 2021, 32, 2111404. [Google Scholar] [CrossRef]
  21. Kondo, Y.; Hino, K.; Kuwahara, Y.; Mori, K.; Yamashita, H. Photosynthesis of hydrogen peroxide from dioxygen and water using aluminium-based metal–organic framework assembled with porphyrin- and pyrene-based linkers. J. Mater. Chem. A 2023, 11, 9530. [Google Scholar] [CrossRef]
  22. Krishnaraj, C.; Sekhar Jena, H.; Bourda, L.; Laemont, A.; Pachfule, P.; Roeser, J.; Chandran, C.V.; Borgmans, S.; Rogge, S.M.J.; Leus, K.; et al. Strongly Reducing (Diarylamino)benzene-Based Covalent Organic Framework for Metal-Free Visible Light Photocatalytic H2O2 Generation. J. Am. Chem. Soc. 2020, 142, 20107–20116. [Google Scholar] [CrossRef] [PubMed]
  23. Qiu, J.; Dai, D.; Yao, J. Tailoring metal–organic frameworks for photocatalytic H2O2 production. Coord. Chem. Rev. 2023, 501, 215597. [Google Scholar] [CrossRef]
  24. Shiraishi, Y.; Kanazawa, S.; Sugano, Y.; Tsukamoto, D.; Sakamoto, H.; Ichikawa, S.; Hirai, T. Highly selective production of hydrogen peroxide on graphitic carbon nitride (g-C3N4) photocatalyst activated by visible light. ACS Catal. 2014, 4, 774–780. [Google Scholar] [CrossRef]
  25. Cote, A.P.; Benin, A.I.; Ockwig, N.W.; O’Keeffe, M.; Matzger, A.J.; Yaghi, O.M. Porous, crystalline, covalent organic frameworks. Science 2005, 310, 1166–1170. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, C.; Zhang, Z.; Zhu, Y.; Yang, C.; Wu, J.; Hu, W. 2D covalent organic frameworks: From synthetic strategies to advanced optical-electrical-magnetic functionalities. Adv. Mater. 2022, 34, 2102290. [Google Scholar] [CrossRef] [PubMed]
  27. Shiraishi, Y.; Takii, T.; Hagi, T.; Mori, S.; Kofuji, Y.; Kitagawa, Y.; Tanaka, S.; Ichikawa, S.; Hirai, T. Resorcinol–formaldehyde resins as metal-free semiconductor photocatalysts for solar-to-hydrogen peroxide energy conversion. Nat. Mater. 2019, 18, 985–993. [Google Scholar] [CrossRef] [PubMed]
  28. Zhou, Z.; Sun, M.; Zhu, Y.; Li, P.; Zhang, Y.; Wang, M.; Shen, Y. A thioether-decorated triazine-based covalent organic framework towards overall H2O2 photosynthesis without sacrificial agents. Appl. Catal. B Environ. 2023, 334, 122862. [Google Scholar] [CrossRef]
  29. Luo, Y.; Zhang, B.; Liu, C.; Xia, D.; Ou, X.; Cai, Y.; Zhou, Y.; Jiang, J.; Han, B. Sulfone-Modified Covalent Organic Frameworks Enabling Efficient Photocatalytic Hydrogen Peroxide Generation via One-Step Two-Electron O2 Reduction. Angew. Chem. Int. Ed. 2023, 62, e202305355. [Google Scholar] [CrossRef]
  30. Liao, Q.; Sun, Q.; Xu, H.; Wang, Y.; Xu, Y.; Li, Z.; Hu, J.; Wang, D.; Li, H.; Xi, K. Regulating Relative Nitrogen Locations of Diazine Functionalized Covalent Organic Frameworks for Overall H2O2 Photosynthesis. Angew. Chem. Int. Ed. 2023, 62, e202310556. [Google Scholar] [CrossRef]
  31. Chen, L.; Wang, L.; Wan, Y.; Zhang, Y.; Qi, Z.; Wu, X.; Xu, H. Acetylene and Diacetylene Functionalized Covalent Triazine Frameworks as Metal-Free Photocatalysts for Hydrogen Peroxide Production: A New Two-Electron Water Oxidation Pathway. Adv. Mater. 2019, 32, 1904433. [Google Scholar] [CrossRef] [PubMed]
  32. Chang, J.-N.; Li, Q.; Shi, J.-W.; Zhang, M.; Zhang, L.; Li, S.; Chen, Y.; Li, S.-L.; Lan, Y.-Q. Oxidation-Reduction Molecular Junction Covalent Organic Frameworks for Full Reaction Photosynthesis of H2O2. Angew. Chem. Int. Ed. 2023, 62, e202218868. [Google Scholar] [CrossRef] [PubMed]
  33. Yue, J.-Y.; Song, L.-P.; Fan, Y.-F.; Pan, Z.-X.; Yang, P.; Ma, Y.; Xu, Q.; Tang, B. Thiophene-Containing Covalent Organic Frameworks for Overall Photocatalytic H2O2 Synthesis in Water and Seawater. Angew. Chem. Int. Ed. 2023, 62, e202309624. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, Y.X.; Li, L.; Tan, H.; Ye, N.; Gu, Y.; Zhao, S.Q.; Zhang, S.P.; Luo, M.C.; Guo, S.J. Fluorination of Covalent Organic Framework Reinforcing the Confinement of Pd Nanoclusters Enhances Hydrogen Peroxide Photosynthesis. J. Am. Chem. Soc. 2023, 145, 19877–19884. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, D.; Chen, W.; Wu, Y.; Wang, L.; Wu, X.; Xu, H.; Chen, L. Covalent Organic Frameworks Containing Dual O2 Reduction Centers for Overall Photosynthetic Hydrogen Peroxide Production. Angew. Chem. Int. Ed. 2022, 62, e202217479. [Google Scholar] [CrossRef]
  36. Qin, C.; Wu, X.; Tang, L.; Chen, X.; Li, M.; Mou, Y.; Su, B.; Wang, S.; Feng, C.; Liu, J.; et al. Dual donor-acceptor covalent organic frameworks for hydrogen peroxide photosynthesis. Nat. Commun. 2023, 14, 5238. [Google Scholar] [CrossRef] [PubMed]
  37. Li, X.; Chen, D.; Li, N.; Xu, Q.; Li, H.; Lu, J. Built-in electric field and oxygen absorption synergistically optimized an organic/inorganic heterojunction for high-efficiency photocatalytic hydrogen peroxide production. J. Colloid Interface Sci. 2023, 648, 664–673. [Google Scholar] [CrossRef]
  38. Wu, M.; Shan, Z.; Wang, J.; Liu, T.; Zhang, G. Three-dimensional covalent organic framework with tty topology for enhanced photocatalytic hydrogen peroxide production. Chem. Eng. J. 2023, 454, 140121. [Google Scholar] [CrossRef]
  39. Isaka, Y.; Kondo, Y.; Kawase, Y.; Kuwahara, Y.; Mori, K.; Yamashita, H. Photocatalytic production of hydrogen peroxide through selective two-electron reduction of dioxygen utilizing amine-functionalized MIL-125 deposited with nickel oxide nanoparticles†. Chem. Commun. 2018, 54, 9270–9273. [Google Scholar] [CrossRef]
  40. Isaka, Y.; Kawase, Y.; Kuwahara, Y.; Mori, K.; Yamashita, H. Two-Phase System Utilizing Hydrophobic Metal–Organic Frameworks (MOFs) for Photocatalytic Synthesis of Hydrogen Peroxide. Angew. Chem. Int. Ed. 2019, 58, 5402–5406. [Google Scholar] [CrossRef]
  41. Yang, Y.; Cheng, B.; Yu, J.; Wang, L.; Ho, W. TiO2/In2S3 S-scheme photocatalyst with enhanced H2O2-production activity. Nano Res. 2021, 16, 4506–4514. [Google Scholar] [CrossRef]
  42. Zhao, Y.; Liu, Y.; Cao, J.; Wang, H.; Shao, M.; Huang, H.; Liu, Y.; Kang, Z. Efficient production of H2O2 via two-channel pathway over ZIF-8/C3N4 composite photocatalyst without any sacrificial agent. Appl. Catal. B Environ. 2020, 278, 119289. [Google Scholar] [CrossRef]
  43. Lee, J.H.; Cho, H.; Park, S.O.; Hwang, J.M.; Hong, Y.; Sharma, P.; Jeon, W.C.; Cho, Y.; Yang, C.; Kwak, S.K.; et al. High performance H2O2 production achieved by sulfur-doped carbon on CdS photocatalyst via inhibiting reverse H2O2 decomposition. Appl. Catal. B Environ. 2021, 284, 119690. [Google Scholar] [CrossRef]
  44. He, T.; Zhao, Y. Covalent Organic Frameworks for Energy Conversion in Photocatalysis. Angew. Chem. Int. Ed. 2023, 62, e202303086. [Google Scholar] [CrossRef]
  45. Yong, Z.; Ma, T. Solar-to-H2O2 Catalyzed by Covalent Organic Frameworks. Angew. Chem. Int. Ed. 2023, 62, e202308980. [Google Scholar] [CrossRef]
  46. Freese, T.; Meijer, J.T.; Feringa, B.L.; Beil, S.B. An organic perspective on photocatalytic production of hydrogen peroxide. Nat. Catal. 2023, 6, 553–558. [Google Scholar] [CrossRef]
  47. Wu, S.; Quan, X. Design Principles and Strategies of Photocatalytic H2O2 Production from O2 Reduction. ACS ES&T Eng. 2022, 2, 1068–1079. [Google Scholar] [CrossRef]
  48. Sheng, B.; Deng, C.; Li, Y.; Xie, S.; Wang, Z.; Sheng, H.; Zhao, J. In Situ Hydroxylation of a Single-Atom Iron Catalyst for Preferential 1O2 Production from H2O2. ACS Catal. 2022, 12, 14679–14688. [Google Scholar] [CrossRef]
  49. Li, L.; Xu, L.; Hu, Z.; Yu, J.C. Enhanced Mass Transfer of Oxygen through a Gas–Liquid–Solid Interface for Photocatalytic Hydrogen Peroxide Production. Adv. Funct. Mater. 2021, 31, 2106120. [Google Scholar] [CrossRef]
  50. Tan, D.; Zhuang, R.; Chen, R.; Ban, M.; Feng, W.; Xu, F.; Chen, X.; Wang, Q. Covalent Organic Frameworks Enable Sustainable Solar to Hydrogen Peroxide. Adv. Funct. Mater. 2024, 34, 2311655. [Google Scholar] [CrossRef]
  51. Li, L.; Lv, X.; Xue, Y.; Shao, H.; Zheng, G.; Han, Q. Custom-Design of Strong Electron/Proton Extractor on COFs for Efficient Photocatalytic H2O2 Production. Angew. Chem. Int. Ed. 2024, 63, e202320218. [Google Scholar] [CrossRef] [PubMed]
  52. Hu, H.; Tao, Y.; Wang, D.; Li, C.; Jiang, Q.; Shi, Y.; Wang, J.; Qin, J.; Zhou, S.; Kong, Y. Rational modification of hydroxy-functionalized covalent organic frameworks for enhanced photocatalytic hydrogen peroxide evolution. J. Colloid Interface Sci. 2022, 629, 750–762. [Google Scholar] [CrossRef] [PubMed]
  53. Tang, X.; Yang, Y.; Li, X.; Wang, X.; Guo, D.; Zhang, S.; Zhang, K.; Wu, J.; Zheng, J.; Zheng, S.; et al. Postmodification of an Amine-Functionalized Covalent Organic Framework for Enantioselective Adsorption of Tyrosine. ACS Appl. Mater. Interfaces 2023, 15, 24836–24845. [Google Scholar] [CrossRef] [PubMed]
  54. Pachfule, P.; Acharjya, A.; Roeser, J.; Langenhahn, T.; Schwarze, M.; Schomäcker, R.; Thomas, A.; Schmidt, J. Diacetylene Functionalized Covalent Organic Framework (COF) for Photocatalytic Hydrogen Generation. J. Am. Chem. Soc. 2018, 140, 1423–1427. [Google Scholar] [CrossRef]
  55. Hou, Y.; Zhou, P.; Liu, F.; Lu, Y.; Tan, H.; Li, Z.; Tong, M.; Ni, J. Efficient Photosynthesis of Hydrogen Peroxide by Cyano-Containing Covalent Organic Frameworks from Water, Air and Sunlight. Angew. Chem. Int. Ed. 2024, 63, e202318562. [Google Scholar] [CrossRef] [PubMed]
  56. Cheng, H.; Lv, H.; Cheng, J.; Wang, L.; Wu, X.; Xu, H. Rational Design of Covalent Heptazine Frameworks with Spatially Separated Redox Centers for High-Efficiency Photocatalytic Hydrogen Peroxide Production. Adv. Mater. 2021, 34, 2107480. [Google Scholar] [CrossRef]
  57. Zhai, L.; Xie, Z.; Cui, C.-X.; Yang, X.; Xu, Q.; Ke, X.; Liu, M.; Qu, L.-B.; Chen, X.; Mi, L. Constructing Synergistic Triazine and Acetylene Cores in Fully Conjugated Covalent Organic Frameworks for Cascade Photocatalytic H2O2 Production. Chem. Mater. 2022, 34, 5232–5240. [Google Scholar] [CrossRef]
  58. Chai, S.; Chen, X.; Zhang, X.; Fang, Y.; Sprick, R.S.; Chen, X. Rational design of covalent organic frameworks for efficient photocatalytic hydrogen peroxide production. Environ. Sci. Nano 2022, 9, 2464. [Google Scholar] [CrossRef]
  59. Shang, Q.; Liu, Y.; Ai, J.; Yan, Y.; Yang, X.; Wang, D.; Liao, G. Embedding Au nanoclusters into the pores of carboxylated COF for the efficient photocatalytic production of hydrogen peroxide. J. Mater. Chem. A 2023, 11, 21109–21122. [Google Scholar] [CrossRef]
  60. Chen, H.; Gao, S.; Huang, G.; Chen, Q.; Gao, Y.; Bi, J. Built-in electric field mediated S-scheme charge migration in COF/In2S3 heterojunction for boosting H2O2 photosynthesis and sterilization. Appl. Catal. B Environ. 2024, 343, 123545. [Google Scholar] [CrossRef]
  61. Feng, Y.; Li, J.; Ye, S.; Gao, S.; Cao, R. Growing COFs in situ on CdS nanorods as core–shell heterojunctions to improve the charge separation efficiency. Sustain. Energy Fuels 2022, 6, 5089. [Google Scholar] [CrossRef]
  62. Zhang, Y.; Hu, Z.; Zhang, H.; Li, H.; Yang, S. Uncovering Original Z Scheme Heterojunctions of COF/MOx (M = Ti, Zn, Zr, Sn, Ce, and Nb) with Ascendant Photocatalytic Selectivity for Virtually 99.9% NO-to-NO3− Oxidation. Adv. Funct. Mater. 2023, 33, 2303851. [Google Scholar] [CrossRef]
  63. Zhang, Y.; Qiu, J.; Zhu, B.; Fedin, M.V.; Cheng, B.; Yu, J.; Zhang, L. ZnO/COF S-scheme heterojunction for improved photocatalytic H2O2 production performance. Chem. Eng. J. 2022, 444, 136584. [Google Scholar] [CrossRef]
  64. Guan, Q.; Zhou, L.-L.; Dong, Y.-B. Construction of covalent organic frameworks via multicomponent reactions. J. Am. Chem. Soc. 2023, 145, 1475–1496. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, J.; Yu, Y.; Cui, J.; Li, X.; Zhang, Y.; Wang, C.; Yu, X.; Ye, J. Defective g-C3N4/covalent organic framework van der Waals heterojunction toward highly efficient S-scheme CO2 photoreduction. Appl. Catal. B Environ. 2021, 301, 120814. [Google Scholar] [CrossRef]
  66. Cong, Y.; Li, X.; Zhang, S.; Zheng, Q.; Zhang, Y.; Lv, S.-W. Embedding carbon quantum dots into crystalline polyimide covalent organic frameworks to enhance water oxidation for achieving dual-channel photocatalytic H2O2 Generation in a wide pH range. ACS Appl. Mater. Interfaces 2023, 15, 43799–43809. [Google Scholar] [CrossRef] [PubMed]
  67. He, Y.; Zhao, J.; Sham, Y.-T.; Gao, S.; Pan, M.; Chen, Q.; Huang, G.; Wong, P.K.; Bi, J. Efficient hydrogen peroxide photosynthesis over CdS/COF for water disinfection: The S-Scheme pathway, oxygen adsorption, and reactor design. ACS Sustain. Chem. Eng. 2023, 11, 17552–17563. [Google Scholar] [CrossRef]
  68. Zhang, H.; Liu, J.; Zhang, Y.; Cheng, B.; Zhu, B.; Wang, L. BiOBr/COF S-scheme photocatalyst for H2O2 production via concerted two-electron pathway. J. Mater. Sci. Technol. 2023, 166, 241–249. [Google Scholar] [CrossRef]
  69. Li, Y.; Chen, W.; Xing, G.; Jiang, D.; Chen, L. New synthetic strategies toward covalent organic frameworks. Chem. Soc. Rev. 2020, 49, 2845–3234. [Google Scholar] [CrossRef]
  70. Campbell, N.L.; Clowes, R.; Ritchie, L.K.; Cooper, A.I. Rapid Microwave Synthesis and Purification of Porous Covalent Organic Frameworks. Chem. Mater. 2009, 21, 204–206. [Google Scholar] [CrossRef]
  71. Yue, J.-Y.; Song, L.-P.; Pan, Z.-X.; Yang, P.; Ma, Y.; Xu, Q.; Tang, B. Regulating the H2O2 Photosynthetic Activity of Covalent Organic Frameworks through Linkage Orientation. ACS Catal. 2024, 14, 4728–4737. [Google Scholar] [CrossRef]
  72. Yang, T.; Zhang, D.; Kong, A.; Zou, Y.; Yuan, L.; Liu, C.; Luo, S.; Wei, G.; Yu, C. Robust Covalent Organic Framework Photocatalysts for H2O2 Production: Linkage Position Matters. Angew. Chem. Int. Ed. 2024, 63, e202404077. [Google Scholar] [CrossRef] [PubMed]
  73. Wang, X.; Jin, Y.; Li, N.; Zhang, H.; Liu, X.; Yang, X.; Pan, H.; Wang, T.; Wang, K.; Qi, D.; et al. 12 Connecting Sites Linked Three-dimensional Covalent Organic Frameworks with Intrinsic Non-interpenetrated shp Topology for Photocatalytic H2O2 Synthesis. Angew. Chem. Int. Ed. 2024, 63, e202401014. [Google Scholar] [CrossRef] [PubMed]
  74. Guan, X.; Fang, Q.; Yan, Y.; Qiu, S. Functional Regulation and Stability Engineering of Three-Dimensional Covalent Organic Frameworks. Acc. Chem. Res. 2022, 55, 1912–1927. [Google Scholar] [CrossRef] [PubMed]
  75. Gui, B.; Lin, G.; Ding, H.; Gao, C.; Mal, A.; Wang, C. Three-Dimensional Covalent Organic Frameworks: From Topology Design to Applications. Acc. Chem. Res. 2020, 53, 2225–2234. [Google Scholar] [CrossRef] [PubMed]
  76. Meng, Y.; Luo, Y.; Shi, J.-L.; Ding, H.; Lang, X.; Chen, W.; Zheng, A.; Sun, J.; Wang, C. 2D and 3D Porphyrinic Covalent Organic Frameworks: The Influence of Dimensionality on Functionality. Angew. Chem. Int. Ed. 2020, 59, 3624–3629. [Google Scholar] [CrossRef]
  77. Chen, L.; Li, S.; Yang, Z.; Chen, C.; Chu, C.; Chen, B. Enhanced photocatalytic hydrogen peroxide production at a solid-liquid-air interface via microenvironment engineering. Appl. Catal. B Environ. 2022, 305, 121066. [Google Scholar] [CrossRef]
  78. Wu, W.; Li, Z.; Liu, S.; Zhang, D.; Cai, B.; Liang, Y.; Wu, M.; Liao, Y.; Zhao, X. Pyridine-based Covalent Organic Frameworks with Pyridyl-Imine Structures for Boosting Photocatalytic H2O2 Production via One-Step 2e Oxygen Reduction. Angew. Chem. Int. Ed. 2024, 63, e202404563. [Google Scholar] [CrossRef]
  79. Xu, C.; Zou, R.; Peng, Y.; Liu, Q.; Ruan, S.; Hu, J. In situ transmission electron microscope studies on one-dimensional nanomaterials: Manipulation, properties and applications. Prog. Mater. Sci. 2020, 113, 100674. [Google Scholar] [CrossRef]
  80. Wang, X.; Zhang, L.; He, S.; Chen, X.; Huang, X.; Zhou, H. Dynamic Imine Exchange Reactions for Facile Synthesis of Imine-Linked Covalent Organic Frameworks. Chem. Mater. 2023, 35, 10070–10077. [Google Scholar] [CrossRef]
  81. Wang, X.; Sun, Y.; Xiao, Y.; Chen, X.; Huang, X.; Zhou, H. Facile Solution-Refluxing Synthesis and Photocatalytic Dye Degradation of a Dynamic Covalent Organic Framework. Molecules 2022, 27, 8002. [Google Scholar] [CrossRef] [PubMed]
  82. Liu, F.; Zhou, P.; Hou, Y.; Tan, H.; Liang, Y.; Liang, J.; Zhang, Q.; Guo, S.; Tong, M.; Ni, J. Covalent organic frameworks for direct photosynthesis of hydrogen peroxide from water, air and sunlight. Nat. Commun. 2023, 14, 4344. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, X.; Cheng, S.; Chen, C.; Wen, X.; Miao, J.; Zhou, B.; Long, M.; Zhang, L. Keto-anthraquinone covalent organic framework for H2O2 photosynthesis with oxygen and alkaline water. Nat. Commun. 2024, 15, 2649. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of photocatalytic H2O2 synthesis mechanism and pathways of water oxidation and oxygen reduction reaction.
Figure 1. Schematic diagram of photocatalytic H2O2 synthesis mechanism and pathways of water oxidation and oxygen reduction reaction.
Catalysts 14 00429 g001
Figure 2. (a) Schematic chemical structures of DAzCOFs with different relative nitrogen positions; (b) EPR spectra of TpDz photocatalytic system after 5 min of dark and visible irradiation; (c) Koutecky–Levich plot acquired from RDE measurements; (d) Comparison of hydrogen peroxide production performance of DAzCOFs with other COF-based photocatalysts [30]. Copyright 2023 Wiley-VCH.
Figure 2. (a) Schematic chemical structures of DAzCOFs with different relative nitrogen positions; (b) EPR spectra of TpDz photocatalytic system after 5 min of dark and visible irradiation; (c) Koutecky–Levich plot acquired from RDE measurements; (d) Comparison of hydrogen peroxide production performance of DAzCOFs with other COF-based photocatalysts [30]. Copyright 2023 Wiley-VCH.
Catalysts 14 00429 g002
Figure 3. (a) Schematic diagram of the synthesis of TD-COF and TT-COF. Performance of TD-COF and TT-COF for photocatalytic production of hydrogen peroxide in water and seawater under (b) O2, (c) air, and (d) Ar atmospheres [33]. Copyright 2023 Wiley-VCH.
Figure 3. (a) Schematic diagram of the synthesis of TD-COF and TT-COF. Performance of TD-COF and TT-COF for photocatalytic production of hydrogen peroxide in water and seawater under (b) O2, (c) air, and (d) Ar atmospheres [33]. Copyright 2023 Wiley-VCH.
Catalysts 14 00429 g003
Figure 4. (a) Schematic synthetic pathways of C-COFs, S-COFs, and FS-COFs; (b) mechanistic schematic of the key steps in the production of hydrogen peroxide by FS-COFs (I: O2 adsorption, II: photogenerated carrier separation, III: proton coupling, IV: H2O2 desorption); (c) comparative plots of metal-free photocatalysts for photocatalytic hydrogen peroxide production without sacrificial agents [29]. Copyright 2023 Wiley-VCH. (d) Schematic representation of palladium metal-isolated clusters interacting with TAPT-PBA COFs and TAPT-TFPA COFs; (e) photocatalytic stability for the production of H2O2 using TAPT-TFPA COFs@Pd-ICs [34]. Copyright 2023 American Chemical Society.
Figure 4. (a) Schematic synthetic pathways of C-COFs, S-COFs, and FS-COFs; (b) mechanistic schematic of the key steps in the production of hydrogen peroxide by FS-COFs (I: O2 adsorption, II: photogenerated carrier separation, III: proton coupling, IV: H2O2 desorption); (c) comparative plots of metal-free photocatalysts for photocatalytic hydrogen peroxide production without sacrificial agents [29]. Copyright 2023 Wiley-VCH. (d) Schematic representation of palladium metal-isolated clusters interacting with TAPT-PBA COFs and TAPT-TFPA COFs; (e) photocatalytic stability for the production of H2O2 using TAPT-TFPA COFs@Pd-ICs [34]. Copyright 2023 American Chemical Society.
Catalysts 14 00429 g004
Figure 5. (a) Schematic diagram of the molecular engineering strategy for the photocatalytic hydrogen peroxide evolution of COFs with isolated redox centers. (b) Schematic route of synthesis of HEP-TAPT-COF and HEP-TAPB-COF [35]. Copyright 2022 Wiley-VCH.
Figure 5. (a) Schematic diagram of the molecular engineering strategy for the photocatalytic hydrogen peroxide evolution of COFs with isolated redox centers. (b) Schematic route of synthesis of HEP-TAPT-COF and HEP-TAPB-COF [35]. Copyright 2022 Wiley-VCH.
Catalysts 14 00429 g005
Figure 6. (a) Schematic route of synthesis of TpaBtt, TapbBtt, and TaptBtt. (b) Schematic illustrations of the direction of electron transfer between functional motifs and imine bonds in TpaBtt, TapbBtt, and TaptBtt. The yellow dashed line represents the motifs and the red dashed line represents the imine bond. Shades of color represent differences in energy [36]. Copyright 2023 Springer Nature.
Figure 6. (a) Schematic route of synthesis of TpaBtt, TapbBtt, and TaptBtt. (b) Schematic illustrations of the direction of electron transfer between functional motifs and imine bonds in TpaBtt, TapbBtt, and TaptBtt. The yellow dashed line represents the motifs and the red dashed line represents the imine bond. Shades of color represent differences in energy [36]. Copyright 2023 Springer Nature.
Catalysts 14 00429 g006
Figure 7. (a) Schematic route of CdS/TpBpy synthesis; (bd) charge-transfer mechanism of CdS/TpBpy; (e) photocatalytic H2O2 synthesis of CdS, TpBpy, and CdS/TpBpy; (f,g) Kf and Kd values for photocatalytic H2O2 decomposition of CdS, TpBpy, and CdS/TpBpy [37]. Copyright 2023 Elsevier.
Figure 7. (a) Schematic route of CdS/TpBpy synthesis; (bd) charge-transfer mechanism of CdS/TpBpy; (e) photocatalytic H2O2 synthesis of CdS, TpBpy, and CdS/TpBpy; (f,g) Kf and Kd values for photocatalytic H2O2 decomposition of CdS, TpBpy, and CdS/TpBpy [37]. Copyright 2023 Elsevier.
Catalysts 14 00429 g007
Figure 8. (a) Schematic synthetic routes of COF-NUST-16 with extended [8 + 4] Tty network and (b) COF-NUST-17 with extended [4 + 2] bex network. Different colors represent different building blocks, purple, orange, and blue represent FBTA-8CHO, TAPPy, and 2D-2F-CHO [38]. Copyright 2022 Elsevier.
Figure 8. (a) Schematic synthetic routes of COF-NUST-16 with extended [8 + 4] Tty network and (b) COF-NUST-17 with extended [4 + 2] bex network. Different colors represent different building blocks, purple, orange, and blue represent FBTA-8CHO, TAPPy, and 2D-2F-CHO [38]. Copyright 2022 Elsevier.
Catalysts 14 00429 g008
Table 1. The photocatalyst for hydrogen peroxide production.
Table 1. The photocatalyst for hydrogen peroxide production.
PhotocatalystsIrradiation ConditionCatalysts AmountReaction
Solution
Reaction
Pathway
H2O2
Generation Rate
SCC
Efficiency (%)
AQY (%)Ref.
TDB-COFSimulated AM 1.5 G illumination10 mg10 mL waterTwo-step 1e ORR and two-step 1e WOR723.5 μmol h−1 g−10.39%1.4%[28]
FS-COFs300 W Xe lamp (λ > 420 nm)5 mg20 mL water2e ORR3904.2 μmol h−1 g−1 6.21% [29]
TpDz-COF300 W Xe lamp (λ > 420 nm)3 mg18 mL water2e ORR7327 μmol h−1 g−10.62%11.9%[30]
CTF-EDDBN300 W Xe lamp (λ > 420 nm)30 mg50 mL water2e ORR and 2e WOR54 μmol h−1 g−10.07% [31]
CTF-BDDBN300 W Xe lamp (λ > 420 nm)30 mg50 mL water2e ORR and 2e WOR97 μmol h−1 g−10.14% [31]
TTF-BT-COF300 W Xe lamp (λ > 420 nm)5 mg10 mL waterTwo-step 1e ORR and two-step 1e WOR2760 μmol h−1 g−10.49%11.19%[32]
TD-COFLED (400–700 nm, 100 mW cm−2)1 mg4 mL seawaterTwo-step 1e ORR and 2e WOR3364 μmol h−1 g−10.15% [33]
TT-COFLED (400–700 nm, 100 mW cm−2)1 mg4 mL seawaterTwo-step 1e ORR and 2e WOR2890 μmol h−1 g−10.14% [33]
TAPT-TFPA COFs@Pd-ICs300 W Xe lamp (λ > 400 nm)10 mg20 mL 10%
ethanol water
Two-step 1e ORR2143 μmol h−1 g−10.82%6.5%[34]
HEP-TAPT-COF300 W Xe lamp (λ > 420 nm)50 mg100 mL water2e ORR87.50 μmol h−10.65%15.35%[35]
HEP-TAPB-COF300 W Xe lamp (λ > 420 nm)50 mg100 mL water2e ORR49.50 μmol h−10.38%9.98%[35]
TaptBtt-COF300 W Xe lamp (λ > 420 nm)15 mg50 mL waterTwo-step 1e ORR and 2e WOR1407 μmol h−1 g−10.296%4.6%[36]
CdS/TpBpyCOF300 W Xe lamp (λ > 420 nm)10 mg50 mL waterTwo-step 1e ORR3600 μmol h−1 g−1 13.4%[37]
COF- NUST -16300 W Xe lamp (λ > 420 nm)5 mg50 mL 10%
ethanol water
Two-step 1e ORR and Two-step 1e WOR1081 μmol h−1 g−1 [38]
MIL-125-NH2500 W Xe lamp (λ > 420 nm)5 mgAcetonitrile/water (4:1) 5 mLTwo-step 1e ORR8 mM h−1 [39]
MIL-125-R7visible light (λ > 420 nm)5 mgWater/benzyl alcohol (2:5) 7 mLTwo-step 1e ORRabout 400 μM h−1 [40]
TiO2/In2S3300 W Xe lamp20 mg40 mL 10%
ethanol water
Two-step 1e ORR376 μmol h−1 L−1 3.42% [41]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Meng, J.; Huang, Y.; Wang, X.; Liao, Y.; Zhang, H.; Dai, W. Photocatalytic Production of Hydrogen Peroxide from Covalent-Organic-Framework-Based Materials: A Mini-Review. Catalysts 2024, 14, 429. https://doi.org/10.3390/catal14070429

AMA Style

Meng J, Huang Y, Wang X, Liao Y, Zhang H, Dai W. Photocatalytic Production of Hydrogen Peroxide from Covalent-Organic-Framework-Based Materials: A Mini-Review. Catalysts. 2024; 14(7):429. https://doi.org/10.3390/catal14070429

Chicago/Turabian Style

Meng, Jiayi, Yamei Huang, Xinglin Wang, Yifan Liao, Huihui Zhang, and Weilin Dai. 2024. "Photocatalytic Production of Hydrogen Peroxide from Covalent-Organic-Framework-Based Materials: A Mini-Review" Catalysts 14, no. 7: 429. https://doi.org/10.3390/catal14070429

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop