Next Article in Journal
Modeling the Impact of Dye Concentration on Polymer Optical Properties via the Complex Refractive Index: A Pathway to Optical Engineering
Next Article in Special Issue
Fabrication, Structural Characterization, and Photon Attenuation Efficiency Investigation of Polymer-Based Composites
Previous Article in Journal
Additive and Lithographic Manufacturing of Biomedical Scaffold Structures Using a Versatile Thiol-Ene Photocurable Resin
Previous Article in Special Issue
Mechanical Properties, Radiation Resistance Performances, and Mechanism Insights of Nitrile Butadiene Rubber Irradiated with High-Dose Gamma Rays
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

COF-Based Photocatalysts for Enhanced Synthesis of Hydrogen Peroxide

by
Deming Tan
1,* and
Xuelin Fan
2
1
School of Mechanical Engineering, Chengdu University, Chengdu 610106, China
2
Institute for Advanced Study, Chengdu University, Chengdu 610106, China
*
Author to whom correspondence should be addressed.
Polymers 2024, 16(5), 659; https://doi.org/10.3390/polym16050659
Submission received: 9 November 2023 / Revised: 4 December 2023 / Accepted: 4 December 2023 / Published: 29 February 2024
(This article belongs to the Special Issue Polymer Materials for Energy, Environment and Radiation Shielding)

Abstract

:
Covalent Organic Frameworks (COFs), with their intrinsic structural regularity and modifiable chemical functionality, have burgeoned as a pivotal material in the realm of photocatalytic hydrogen peroxide (H2O2) synthesis. This article reviews the recent advancements and multifaceted approaches employed in using the unique properties of COFs for high-efficient photocatalytic H2O2 production. We first introduced COFs and their advantages in the photocatalytic synthesis of H2O2. Subsequently, we spotlight the principles and evaluation of photocatalytic H2O2 generation, followed by various strategies for the incorporation of active sites aiming to optimize the separation and transfer of photoinduced charge carriers. Finally, we explore the challenges and future prospects, emphasizing the necessity for a deeper mechanistic understanding and the development of scalable and economically viable COF-based photocatalysts for sustainable H2O2 production.

1. Introduction

The quest for sustainable and clean energy sources has become a paramount concern in the face of escalating global energy demands and the pressing challenges of climate change. Solar energy, being abundant and renewable, stands out as a promising candidate to address these challenges. Efficiently harnessing the vast potential of solar energy not only offers a solution to the impending energy crisis but also holds the promise of reducing the carbon footprint associated with conventional fossil fuels. Photocatalysis is a process that uses light to drive chemical reactions and has emerged as an important technology [1,2,3,4,5,6]. By converting solar energy into chemical energy, photocatalysis offers a dual advantage [4,7]: it provides an avenue for sustainable energy storage and paves the way for the synthesis of valuable chemicals, including H2, CO, and hydrogen peroxide (H2O2).
Among various chemicals synthesized by photosynthesis, H2O2 has attracted significant attention [2,8,9,10]. H2O2 is valuable for its multifaceted applications and environmentally amicable nature, and has become an indispensable chemical field like industrial processes, environmental remediation, and healthcare and medical applications [11,12,13,14]. Its ability to decompose into water and oxygen underpins its appeal as an eco-friendly oxidant, minimizing the risk of generating secondary pollutants. The conventional anthraquinone oxidation process [15], which has been the industrial applicable for H2O2 production, is increasingly being scrutinized for its inherent drawbacks. These include not only the substantial energy consumption and the utilization of hazardous substrates but also the generation of significant amounts of waste, which poses considerable environmental and economic challenges. In light of these limitations, the researchers have pivoted towards seeking alternative, sustainable, and cleaner methods for H2O2 production, with a particular emphasis on minimizing environmental repercussions. Photocatalytic synthesis of H2O2 has emerged as a compelling alternative, offering the prospect of harnessing solar energy directly to drive chemical reactions, thereby circumventing the need for energy-intensive processes and deleterious chemicals [16,17]. This approach not only aligns with the global shift towards sustainable energy but also presents a pathway for the localized, on-demand production of H2O2, reducing the need for storage and transportation.
In the realm of photocatalytic H2O2 synthesis, the role of catalysts is paramount, dictating the efficiency, selectivity, and stability of the photocatalytic processes. Conventional photocatalysts are predominantly based on precious metals such as platinum [18], palladium [19], and gold [20,21], demonstrating commendable performance in facilitating the production of H2O2. However, the deployment of these noble metal-based catalysts is significantly hampered by their scarcity in the Earth’s crust, which intrinsically leads to high costs and poses sustainability concerns, especially in the context of large-scale applications and global accessibility. Consequently, the exploration of alternative non-metal-based photocatalysts has become a focal point in contemporary research, aiming to circumvent the limitations associated with noble-metal catalysts. Metal-free photocatalysts, particularly those based on linear polymers [22], polymeric carbon nitride (PCN) [23,24], polymer resins [25], supramolecular coordination [26,27], and covalent organic frameworks (COFs) [2,28,29], have emerged as promising candidates, offering the advantages of abundance and low cost under photocatalytic conditions. Among these, COFs, with their intrinsic porosity, tunable structures, and the ability to incorporate a myriad of organic functional groups (Figure 1), have garnered substantial attention since 2020 [30,31,32].
COFs are typically a class of porous polymers formed by organic building blocks connected through covalent bonds. They were first synthesized by Yaghi et al. under solvothermal conditions through the self-condensation of phenyl diboronic acid (PDBA) and the co-condensation of PDBA with hexahydroxytriphenylene (HHTP) [33]. This work opened the door to COF research. Subsequently, various methods for synthesizing COFs have been reported, including solvothermal [34,35,36,37,38], microwave [39,40,41], ionothermal [42,43,44], and mechanochemical methods [45,46] for powder synthesis, and interfacial methods [47,48,49] for thin-film synthesis. At the same time, a wide variety of organic building blocks and linkages have been reported. To date, reported linkages include boroxine [50], boronate-ester [51,52,53], imine [54,55,56,57], hydrazone [58,59], squaraine [60,61,62], azine [63,64,65], imide [66,67], C=C [68,69,70,71], 1,4-dioxin linkage [72,73], among others. The COFs synthesized by these methods have shown great potential in applications such as sensing [74,75,76], catalysis [5,77,78,79], energy storage and conversion, [6,80,81,82,83] organic electronic devices [84,85,86,87,88,89,90,91,92], etc. [6,80,81,82,83,93,94,95,96].
COFs not only possess ordered pore structure and high surface area but also offer the possibility of fine-tailoring their electronic and structural properties to optimize charge separation and transfer, thereby standing out as a promising class of materials in the ongoing quest for sustainable and efficient photocatalytic processes. Despite the growing number of literature reports on COF photocatalytic synthesis of H2O2 since 2020, there is a scarcity of review papers summarizing the principal advancements in this field. Therefore, this review aims to acquaint readers with the advancement in this nascent field, as well as the prevailing challenges. Finally, perspectives are provided on the future development of this field.

2. Principles of Photocatalytic H2O2 Generation

Equation (1) illustrates the full process of photocatalytic synthesis of H2O2. This procedure encompasses two distinct half-reactions, namely the oxygen reduction reaction (ORR) (Equations (2) and (3)) and the water oxidation reaction (WOR) (Equation (4)). In response to the impetus of photons, the electrons within the catalyst undergo a transition from the valence band (VB) to the conduction band (CB), thus giving rise to the emergence of photo-excited h+ and e species. The subsequent migration of these charges to the catalyst’s surface facilitates their active participation in a cascading series of reduction-oxidation reactions (Figure 2a), thereby selectively yielding H2O2.
The photo-excited e with reducibility is capable of reacting with oxygen through a 2e ORR process for the generation of H2O2 (see Equation (2)). It is worth noting that the oxygen involved in this reaction may be derived from either the 4e WOR or the atmospheric oxygen. Theoretically, as shown in Figure 2b, the 2e ORR can proceed directly through a 2e process (O2 + 2e + 2H+ → H2O2) or two consecutive 1e reactions via the superoxide radical intermediate (·O2) (III of Figure 2b). In parallel, the photo-excited h+, which exhibits oxidizability, enables water molecules to generate H2O2 through a directly one-step 2e WOR (see Equation (3)) or two-step 1e WOR, or to produce O2 through a 4e WOR (see Equation (4)). Similar to the two-step 1e ORR process, in the two-step 1e WOR process, the photo-induced proton h+ can initially oxidize H2O to generate a hydroxyl radical (·OH) intermediate. Subsequently, H2O2 is formed indirectly through the combination of two ·OH (VI of Figure 2b). These intermediates can be identified through in-situ characterization techniques. In-situ diffuse reflectance infrared Fourier transform spectroscopy is extensively utilized for the characterization of intermediates in the photocatalytic generation of H2O2. A well-designed photocatalyst aimed at the overall reaction should possess functional groups that effectively promote both the 2e ORR and 2e WOR. These two reactions, in turn, serve to maintain charge balance by, respectively, consuming e and h+. In instances where catalysts are specifically engaged in one of the half-reactions, sacrificial reagents are often employed to consume uninvolved charges, thereby safeguarding the catalyst and directing the reaction’s selectivity.
2 H 2 O + O 2 h v 2 H 2 O 2
O 2 + 2 e + 2 H + H 2 O 2
2 H 2 O + 2 h + H 2 O 2 + 2 H +
H 2 O + 4 h + O 2 + 4 H +
Nonetheless, it is worth noting that the solar-to-chemical energy conversion efficiency (SCC) in the context of photocatalytic H2O2 synthesis remains relatively low at present, seldom surpassing 1% [97]. This efficiency discrepancy is quite pronounced when compared to the performance observed in photocatalytic hydrogen production from water [98,99]. Several factors contribute to this relatively low efficiency, including limited light absorption, a propensity for charge recombination, and challenges in achieving desirable selectivity towards H2O2 [97,100,101]. Particularly, achieving selectivity towards H2O2 presents a significant hurdle (Figure 2). The WOR process commonly tends to favor O2 production via the 4e pathway rather than H2O2 production through the 2e pathway [17]. Additionally, side reactions and H2O2 decomposition also impact both the yield and selectivity [102,103]. Through the desirable optimization of catalyst structures and reaction conditions, significant enhancements can be achieved in terms of the yield and selectivity of photocatalytic H2O2 synthesis.

3. Evaluation of Photocatalytic H2O2 Generation

The evaluation criteria for photocatalytic H2O2 production mainly include H2O2 production rate, apparent quantum yield (AQY), and solar-to-chemical energy conversion (SCC) efficiency. The rate of H2O2 production is a critical parameter for quantifying the speed of H2O2 formation in photocatalytic processes. This rate is commonly delineated in millimoles per hour per gram (mmol h−1 g−1), referencing the catalyst’s specific mass. AQY quantifies the efficiency of H2O2 generation in relation to photon absorption by the catalyst, which can be calculated by Formula (5). So far, the highest reported AQY value for a COF catalyst is 38% [10]. SCC measures the efficiency of transforming solar energy into chemical energy stored in H2O2 and can be calculated by Formula (6), in which the total input power (W) can be obtained by multiplying the irradiation intensity of the Xenon lamp by the irradiated area and time.
A Q Y % = H 2 O 2 produced mol × 2 photon number entered into the reactor mol × 100 % = [ N a × h × c ] H 2 O 2 produced mol × 2 I × S × t × λ × 100 %
where,
  • Na (Avogadro’s constant) = 6.02 × 1023 mol−1;
  • h (Planck constant) = 6.626 × 10−34 J·s;
  • c (Light speed) = 3 × 108 m/s;
  • S = Irradiation area (cm2) = 11.5 cm2;
  • I = The intensity of irradiation light (W/cm2);
  • t = The photoreaction time (s) = 3600 s;
  • λ = The wavelength of the monochromatic light (nm) = 420 × 10−9.
SCC effeciency % = Δ G for H 2 O 2 generation J mol 1 H 2 O 2 formed mol Total input power W Reaction time s × 100 %
where ΔG is 117 kJ mol−1.

4. COFs for Light-Driven H2O2 Production

In the realm of photocatalytic H2O2 synthesis, COF photocatalysts have demonstrated significant potential for application. These photocatalysts can be primarily categorized into three main types: firstly, those dedicated to catalyzing the 2e ORR; secondly, those focusing on the 2e WOR; and thirdly, those capable of simultaneously catalyzing both 2e ORR and 2e WOR. Compared to the 2e WOR process, the 2e ORR is relatively easier to achieve. As a result, most of the current research on COF-based photocatalytic production of H2O2 is realized through optimizing the 2e ORR process.
To further enhance the catalytic efficiency of the 2e ORR and WOR, researchers often attempt to introduce specific active sites to facilitate the reactions. According to literature, the discovered active sites for the 2e ORR include pyridazine [104], pyridine [105], pyrene [28], bipyridine [106], benzene [107,108], and diarylamine [109], while those for 2e WOR comprise bipyridine [106], triazine [110], diacetylene [111,112], acetylene [111,112], among others. The incorporation of these active sites not only improves the efficiency of COF photocatalysts in H2O2 synthesis but also provides valuable insights for a deeper understanding and optimization of these two crucial reaction processes. From Table 1, it can be observed that the AQY mostly ranges between 1% and 16%, the H2O2 production rate primarily ranges from 100 to 7500 μmol·h−1·g−1, and the SCC is relatively low, with the highest being 0.82%. Through in-depth studies and optimization of active sites, it is anticipated that COF photocatalysts will achieve better performance in the field of photocatalytic H2O2 synthesis. In the following section, we will elaborate in detail on the strategies for introducing active sites.

4.1. Functional Groups Modification

Functional group modification in COF photocatalysts is a pivotal strategy that involves the intricate alteration or introduction of specific chemical groups within the COF structure to enhance their photocatalytic efficiency. Firstly, the tailored functional groups can optimize the electronic properties of COFs, enhancing visible light absorption and improving the charge separation efficiency. Additionally, the modified functional groups can enhance the stability of COFs under various operational conditions, ensuring consistent performance over extended periods. Furthermore, the strategic incorporation or withdrawing groups aids in the efficient transfer of charges, reducing the recombination of electron-hole pairs and amplifying the yield of H2O2. Luo et al. explore the use of sulfone-modified COFs as photocatalysts for H2O2 generation (Figure 3a) [113]. Sulfone units were introduced into the COFs as a strong electron-accepting core. Under LED visible light irradiation, FS-COFs exhibited an H2O2 yield of 1501.6 μmol·h−1·g−1 from pure water (Figure 3b), which was approximately three times more efficient than COFs with a similar structure but without sulfone sites (yield of 487.6 μmol·h−1·g−1). Notably, under Xe light irradiation, FS-COFs achieved an H2O2 yield of 3904.2 μmol·h−1·g−1. Mechanism study reveals that the incorporation of sulfone units into COFs facilitated the direct reduction in O2 to H2O2 via the one-step 2e ORR route. This was attributed to the improved separation of e-h+ pairs, enhanced protonation property, and altered O2 adsorption configuration on FS-COFs. The formation of 1,4-endoperoxide intermediate species favored the direct reduction in O2 to H2O2, avoiding the generation of reactive radicals and promoting higher H2O2 yields.
Similarly, Sun et al. introduced pyrene into COFs and obtained a series of pyrene-based COFs [28]. The performance of pyrene-based COFs was compared with other COFs, and the results showed that the presence of pyrene active sites in a dense concentration in a limited surface area led to undesirable H2O2 decomposition. This finding indicated that the distribution of pyrene units over a large surface area of the COFs played a crucial role in catalytic performance. To address the issue of H2O2 decomposition, a two-phase reaction system (water-benzyl alcohol) was employed. This approach effectively inhibited H2O2 decomposition and allowed for efficient photocatalytic H2O2 generation. Density Functional Theory (DFT) calculations revealed that the pyrene unit was more active for H2O2 production compared to bipyridine and (diarylamino)benzene units. In the presence of the sacrificial agent benzyl alcohol and under illumination with a light source ranging from 420–700 nm, the photocatalytic yield of H2O2 was 1242 μmol·h−1·g−1, with an AQY of 4.5% at 420 nm. In the absence of a sacrificial agent, satisfactory catalytic performance can still be achieved in H2O2 synthesis. Wu and co-workers synthesized thiourea-functionalized Covalent Triazine Framework (Bpt-CTF) (Figure 4a), which demonstrates a photocatalytic H2O2 production rate of 3268.1 μmol·h−1·g−1 (Figure 4b,c) without the necessity for sacrificial agents or cocatalysts [114]. This represents an improvement of over an order of magnitude compared to the unfunctionalized CTF (Dc-CTF), and it exhibits an AQY of 8.6% at 400 nm. Mechanistic studies demonstrate that the introduction of polar (thio)urea moieties onto CTFs leads to polarization enhancement, which significantly promotes charge separation and transfer (Figure 4d,e). This enhanced polarization facilitates efficient proton transfer, ultimately promoting the photosynthesis of H2O2. The triazine units within the CTFs act as the photoreduction site for the 2e ORR to produce H2O2, while the accumulated holes at the thiourea sites accelerate the WOR kinetics [114].
The H2O2 production rate can be further elevated to 7327 μmol·h−1·g−1 [104]. Liao et al. present the development of diazine functionalized COFs (DAzCOFs) for photosynthesis of H2O2 (Figure 5a) [104]. The study focuses on the influence of nitrogen atom positions in N-heterocycles for catalytic activity. The evaluation revealed that pyridazine embedded in TpDz exhibited a remarkable photocatalytic activity for overall H2O2 photosynthesis. The optimized H2O2 production rate of TpDz was measured to be 7327 μmol·h−1·g−1 in pure water (Figure 5b,c), ranking it as one of the best COF-based photocatalysts in photosynthesis of H2O2. Additionally, the SCC efficiency of TpDz was found to be 0.62%, surpassing natural plants. DFT calculations indicate that pyridazine acts as a superior reactive site for stabilizing endoperoxide intermediate species in the direct 2e ORR pathway compared to pyrimidine. Conversely, pyrazine exhibits a propensity to generate ·OOH for the two-step 1e ORR pathway owing to its separated nitrogen atoms.
The research pertaining to the aforementioned functional group modification primarily involves the generation of H2O2 through the ORR process. The production of H2O2 through the 2e WOR remains elusive in the field. However, simultaneously utilizing both ORR and WOR processes for H2O2 production would be particularly appealing, especially in terms of energy conservation and pollution reduction. Chen and co-workers incorporated acetylene and diacetylene into CTFs and successfully obtained two new CTFs, namely CTF-EDDBN and CTF-BDDBN [111]. Additionally, they demonstrate that exfoliated nanosheets derived from CTF-EDDBN and CTF-BDDEN can generate H2O2 through ORR and a new WOR process when subjected to visible light irradiation. The findings indicate that CTF-BDDBN exhibits a BET surface area of 63.5 m2·g−1, a H2O2 production rate of 97.2 μmol·h−1·g−1, and a SCC efficiency of 0.14%. DFT calculation reveals that the introduction of acetylene (−C≡C−) or diacetylene (−C≡C−C≡C−) groups into CTFs has been found to greatly enhance the photocatalytic production of H2O2. This improvement can be attributed to the presence of −C≡C− triple bonds, whose presence plays an important role in modulating the electronic structures of CTFs and suppressing charge recombination. Moreover, the incorporation of acetylene and diacetylene moieties effectively reduces the energy required for the formation of OH* and enables a new pathway for the 2e WOR to produce H2O2.

4.2. Heteroatom Doping

Heteroatom doping refers to the strategy of introducing atoms other than carbon into the framework of COFs to modify their electronic structure and enhance their photocatalytic performance. The heteroatoms, such as nitrogen, boron, or sulfur, can introduce new energy levels within the band gap of COFs, thus facilitating the transfer and separation of photogenerated charges. Moreover, the doped heteroatoms can also serve as active sites for the adsorption and activation of molecular oxygen, which is a key step in the photocatalytic production of H2O2. Additionally, heteroatom doping can also enhance structural stability, which further contributes to their photocatalytic activity. Wang et al. present the development of a fluorinated COF (TF50-COF) photocatalyst for the photosynthesis of H2O2 (Figure 6a) [107]. The evaluation suggests that TF50-COF demonstrated an AQY of 5.1% at 400 nm and a SCC efficiency of 0.17% (Figure 6g). The production of H2O2 was evaluated under different conditions. When irradiated with λ > 400 nm, light, and using ethanol as a sacrificial agent, TF50-COF achieved an H2O2 production rate of 1739 μmol·h−1·g−1 (Figure 6e,f). In comparison, H-COF, a non-fluorinated COF, only produced 516 μmol·h−1·g−1 of H2O2. DFT calculations were performed to comprehend the process of TF50-COF catalytical H2O2 production, indicating that the incorporation of F-substituents resulted in a profusion of Lewis acid sites (Figure 6d). This, in turn, affected the electronic distribution of adjacent carbon structures, resulting in the creation of highly active sites for O2 adsorption. Moreover, this substitution expanded the catalyst’s sensitivity to visible light, simultaneously enhancing the efficiency of charge separation.

4.3. Donor-Acceptor (D-A) Configuration

Enhancing the photocatalytic performance necessitates a crucial stage of fine-tuning the separation and transfer of photogenerated carriers during redox reactions. To achieve this goal, it is necessary to introduce both electron acceptors and electron donors simultaneously. These acceptors and donors are then utilized to construct an innovative framework known as donor-acceptor COFs (D-A COFs). Zhai et al. presented a study on the construction of synergistic triazine and acetylene cores in COFs for photocatalytic H2O2 production (Figure 7a) [115]. The experimental process involves the integration of triazine and acetylene units into COFs, namely EBA-COF and BTEA-COF. These COFs exhibit spatial separation of the triazine and acetylene cores, leading to efficient charge separation and suppressed charge recombination. Moreover, the C≡C linkage within the COFs facilitates the transport of electrons along the frameworks. The findings suggest that the integrated triazine and acetylene units synergistically promote H2O2 synthesis through a 2e ORR pathway. Specifically, the EBA-COF demonstrates an H2O2 production rate of 1830 μmol·h−1·g−1 (Figure 7d,e). Mechanism investigation demonstrated that the electron-deficient triazine moiety with a high content of nitrogen atoms exhibits a high electron affinity. Additionally, the acetylenic groups act as bridges between electron donors and acceptors, facilitating the efficient transfer of photogenerated e [115]. In another investigation, Cheng et al. designed covalent heptazine frameworks (CHF) with separated redox centers, aiming to enhance charge separation and achieve high H2O2 production [112]. The results revealed that the 2e ORR occurred on the heptazine moiety, while the 2e WOR took place on the acetylene or diacetylene bond in the CHF. The exciton binding energy of the diacetylene-containing polymer was found to be 24 meV. Under simulated solar irradiation, the rationally designed CHFs achieved a SCC efficiency of 0.78%. Additionally, mechanism investigation showed that the significant differences in electronegativity between the s-heptazine moiety and different biphenyl compounds facilitated efficient charge transfer from peripheral groups to the central s-heptazine moiety [112].

5. Outlook and Perspective

In the last three years, COFs have emerged as promising photocatalysts for the synthesis of H2O2 due to their intrinsic properties such as tunable porosity, structural diversity, and modifiable electronic structure. The exploration of COFs in this realm has unveiled a spectrum of results, demonstrating enhanced photocatalytic activity and stability under various operational conditions. Notably, the strategic incorporation of active sites and the optimization of electronic properties are important in amplifying the photocatalytic efficiency of COFs for H2O2 production. Currently, reported active sites for the ORR include pyridine, bipyridine, pyridazine, benzene, diarylamine, pyrene, etc., while active sites for the WOR include diacetylene, triazine, acetylene, bipyridine, and others.
Looking forward, the prospects of COFs in photocatalytic H2O2 synthesis appear to be expansive yet challenging. Some challenges still exist in this field. The pore size of most reported COFs for photocatalytic H2O2 production falls within the 1 to 3 nm range. However, there is still a relative lack of research on how the size of these pores affects the photocatalytic process. Additionally, the stability of COFs also restricts their industrial applications. COFs can sometimes suffer from lower stability, especially in aqueous environments, which may limit their practical applications. Compared to inorganic photocatalysts, COFs may not match the robustness and long-term stability of these more established materials. To enhance their stability, more appropriate building blocks and linkage chemistry can be selected or fabricate composites by combining COFs with materials possessing better stabilities. Future research should delve deeper into the mechanistic pathways of photocatalytic reactions of COFs, which is important for rational design and functionalization strategies. Moreover, addressing the scalability and economic viability of COF-based photocatalytic systems is imperative to transition from laboratory-scale research to real-world applications. The exploration of novel COF structures, the incorporation of alternative active sites, and the development of hybrid systems that combine the advantages of various photocatalytic materials will potentially pave the way for breakthroughs in sustainable H2O2 production.

Author Contributions

Writing—original draft, D.T.; Review & editing, D.T. and X.F.; Project administration, D.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Chengdu University Youth Talent Research Startup Fund (30/2081921080) and the project of the National Natural Science Foundation of China (52203133).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yong, Z.; Ma, T. Solar-to-H2O2 Catalyzed by Covalent Organic Frameworks. Angew. Chem. Int. Ed. 2023, 135, e202308980. [Google Scholar] [CrossRef]
  2. Zhang, X.; Zhang, J.; Miao, J.; Wen, X.; Chen, C.; Zhou, B.; Long, M. Keto-enamine-based covalent organic framework with controllable anthraquinone moieties for superior H2O2 photosynthesis from O2 and water. Chem. Eng. J. 2023, 466, 143085. [Google Scholar] [CrossRef]
  3. Yang, T.; Chen, Y.; Wang, Y.; Peng, X.; Kong, A. Weakly Hydrophilic Imine-Linked Covalent Benzene-Acetylene Frameworks for Photocatalytic H2O2 Production in the Two-Phase System. ACS Appl. Mater. Interfaces 2023, 15, 8066–8075. [Google Scholar] [CrossRef] [PubMed]
  4. 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]
  5. Li, J.; Gao, S.Y.; Liu, J.; Ye, S.; Feng, Y.; Si, D.H.; Cao, R. Use in Photoredox Catalysis of Stable Donor–Acceptor Covalent Organic Frameworks and Membrane Strategy. Adv. Funct. Mater. 2023, 33, 2305735. [Google Scholar] [CrossRef]
  6. Chatterjee, A.; Sun, J.; Rawat, K.S.; Van Speybroeck, V.; Van Der Voort, P. Exploring the Charge Storage Dynamics in Donor–Acceptor Covalent Organic Frameworks Based Supercapacitors by Employing Ionic Liquid Electrolyte. Small 2023, 19, 2303189. [Google Scholar] [CrossRef] [PubMed]
  7. Kofuji, Y.; Ohkita, S.; Shiraishi, Y.; Sakamoto, H.; Tanaka, S.; Ichikawa, S.; Hirai, T. Graphitic Carbon Nitride Doped with Biphenyl Diimide: Efficient Photocatalyst for Hydrogen Peroxide Production from Water and Molecular Oxygen by Sunlight. ACS Catal. 2016, 6, 7021–7029. [Google Scholar] [CrossRef]
  8. Yu, F.Y.; Zhou, Y.J.; Tan, H.Q.; Li, Y.G.; Kang, Z.H. Versatile Photoelectrocatalysis Strategy Raising Up the Green Production of Hydrogen Peroxide. Adv. Energy Mater. 2023, 13, 2300119. [Google Scholar] [CrossRef]
  9. Tian, Q.; Zeng, X.K.; Zhao, C.; Jing, L.Y.; Zhang, X.W.; Liu, J. Exceptional Photocatalytic Hydrogen Peroxide Production from Sandwich-Structured Graphene Interlayered Phenolic Resins Nanosheets with Mesoporous Channels. Adv. Funct. Mater. 2023, 33, 2213173. [Google Scholar] [CrossRef]
  10. Shao, C.; He, Q.; Zhang, M.; Jia, L.; Ji, Y.; Hu, Y.; Li, Y.; Huang, W.; Li, Y. A covalent organic framework inspired by C3N4 for photosynthesis of hydrogen peroxide with high quantum efficiency. Chin. J. Catal. 2023, 46, 28–35. [Google Scholar] [CrossRef]
  11. Jones, C.W. Applications of Hydrogen Peroxide and Derivatives; Royal Society of Chemistry: Cambridge, UK, 1999; Volume 2. [Google Scholar]
  12. Targhan, H.; Evans, P.; Bahrami, K.J. A review of the role of hydrogen peroxide in organic transformations. Ind. Eng. Chem. 2021, 104, 295–332. [Google Scholar] [CrossRef]
  13. Urban, M.V.; Rath, T.; Radtke, C. Hydrogen peroxide (H2O2): A review of its use in surgery. Wien. Med. Wochenschr. 2017, 169, 222–225. [Google Scholar] [CrossRef]
  14. Ciriminna, R.; Albanese, L.; Meneguzzo, F.; Pagliaro, M. Hydrogen peroxide: A key chemical for today’s sustainable development. ChemSusChem 2016, 9, 3374–3381. [Google Scholar] [CrossRef]
  15. Campos-Martin, J.M.; Blanco-Brieva, G.; Fierro, J.L. Hydrogen peroxide synthesis: An outlook beyond the anthraquinone process. Angew. Chem. Int. Ed. 2006, 45, 6962–6984. [Google Scholar] [CrossRef] [PubMed]
  16. Yu, W.; Hu, C.; Bai, L.; Tian, N.; Zhang, Y.; Huang, H. Photocatalytic hydrogen peroxide evolution: What is the most effective strategy? Nano Energy 2022, 104, 107906. [Google Scholar] [CrossRef]
  17. Zeng, X.; Liu, Y.; Hu, X.; Zhang, X. Photocatalytic hydrogen peroxide evolution: What is the most effective strategy? Green Chem. 2021, 23, 1466–1494. [Google Scholar] [CrossRef]
  18. Liu, D.; Shen, J.; Xie, Y.; Qiu, C.; Zhang, Z.; Long, J.; Lin, H.; Wang, X. Metallic Pt and PtO2 Dual-Cocatalyst-Loaded Binary Composite RGO-CNx for the Photocatalytic Production of Hydrogen and Hydrogen Peroxide. ACS Sustain. Chem. Eng. 2021, 9, 6380–6389. [Google Scholar] [CrossRef]
  19. Chu, C.; Huang, D.; Zhu, Q.; Stavitski, E.; Spies, J.A.; Pan, Z.; Mao, J.; Xin, H.L.; Schmuttenmaer, C.A.; Hu, S. Electronic tuning of metal nanoparticles for highly efficient photocatalytic hydrogen peroxide production. ACS Catal. 2018, 9, 626–631. [Google Scholar] [CrossRef]
  20. Hirakawa, H.; Shiota, S.; Shiraishi, Y.; Sakamoto, H.; Ichikawa, S.; Hirai, T. Au Nanoparticles Supported on BiVO4: Effective Inorganic Photocatalysts for H2O2 Production from Water and O2 under Visible Light. ACS Catal. 2016, 6, 4976–4982. [Google Scholar] [CrossRef]
  21. Zuo, G.; Liu, S.; Wang, L.; Song, H.; Zong, P.; Hou, W.; Li, B.; Guo, Z.; Meng, X.; Du, Y. Finely dispersed Au nanoparticles on graphitic carbon nitride as highly active photocatalyst for hydrogen peroxide production. Catal. Commun. 2019, 123, 69–72. [Google Scholar] [CrossRef]
  22. Liu, L.; Gao, M.Y.; Yang, H.; Wang, X.; Li, X.; Cooper, A.I. Linear Conjugated Polymers for Solar-Driven Hydrogen Peroxide Production: The Importance of Catalyst Stability. J. Am. Chem. Soc. 2021, 143, 19287–19293. [Google Scholar] [CrossRef]
  23. 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]
  24. Liu, B.; Du, J.; Ke, G.; Jia, B.; Huang, Y.; He, H.; Zhou, Y.; Zou, Z. Boosting O2 Reduction and H2O Dehydrogenation Kinetics: Surface N-Hydroxymethylation of g-C3N4 Photocatalysts for the Efficient Production of H2O2. Adv. Funct. Mater. 2022, 32, 2111125. [Google Scholar] [CrossRef]
  25. Wang, X.; Yang, X.; Zhao, C.; Pi, Y.; Li, X.; Jia, Z.; Zhou, S.; Zhao, J.; Wu, L.; Liu, J. Ambient Preparation of Benzoxazine-based Phenolic Resins Enables Long-term Sustainable Photosynthesis of Hydrogen Peroxide. Angew. Chem. Int. Ed. 2023, 62, e202302829. [Google Scholar] [CrossRef]
  26. Zhang, X.; Jiang, S.; Sun, L.X.; Xing, Y.H.; Bai, F.Y. Synthesis and structure of a 3D supramolecular layered Bi-MOF and its application in photocatalytic degradation of dyes. J. Mol. Struct. 2022, 1270, 133895. [Google Scholar] [CrossRef]
  27. Ham, R.; Nielsen, C.J.; Pullen, S.; Reek, J.N. Supramolecular Coordination Cages for Artificial Photosynthesis and Synthetic Photocatalysis. Chem. Rev. 2023, 123, 5225. [Google Scholar] [CrossRef] [PubMed]
  28. Sun, J.; Sekhar Jena, H.; Krishnaraj, C.; Singh Rawat, K.; Abednatanzi, S.; Chakraborty, J.; Laemont, A.; Liu, W.; Chen, H.; Liu, Y.Y.; et al. Pyrene-Based Covalent Organic Frameworks for Photocatalytic Hydrogen Peroxide Production. Angew. Chem. Int. Ed. 2023, 62, e202216719. [Google Scholar] [CrossRef] [PubMed]
  29. Zhang, Y.; Pan, C.; Bian, G.; Xu, J.; Dong, Y.; Zhang, Y.; Lou, Y.; Liu, W.; Zhu, Y. H2O2 generation from O2 and H2O on a near-infrared absorbing porphyrin supramolecular photocatalyst. Nat. Energy 2023, 8, 361–371. [Google Scholar] [CrossRef]
  30. Gong, Y.-N.; Guan, X.; Jiang, H.-L. Covalent organic frameworks for photocatalysis: Synthesis, structural features, fundamentals and performance. Coord. Chem. Rev. 2023, 475, 214889. [Google Scholar] [CrossRef]
  31. Tan, F.; Zheng, Y.; Zhou, Z.; Wang, H.; Dong, X.; Yang, J.; Ou, Z.; Qi, H.; Liu, W.; Zheng, Z.; et al. Aqueous Synthesis of Covalent Organic Frameworks as Photocatalysts for Hydrogen Peroxide Production. CCS Chem. 2022, 4, 3751–3761. [Google Scholar] [CrossRef]
  32. Li, H.; Wang, L.; Yu, G. Covalent organic frameworks: Design, synthesis, and performance for photocatalytic applications. Nano Today 2021, 40, 101247. [Google Scholar] [CrossRef]
  33. Côté, 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]
  34. Huang, W.; Jiang, Y.; Li, X.; Li, X.; Wang, J.; Wu, Q.; Liu, X. Solvothermal synthesis of microporous, crystalline covalent organic framework nanofibers and their colorimetric nanohybrid structures. ACS Appl. Mater. Interfaces 2013, 5, 8845–8849. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, X.; Huang, N.; Gao, J.; Xu, H.; Xu, F.; Jiang, D. Towards covalent organic frameworks with predesignable and aligned open docking sites. Chem. Commun. 2014, 50, 6161–6163. [Google Scholar] [CrossRef] [PubMed]
  36. Xu, L.; Ding, S.-Y.; Liu, J.; Sun, J.; Wang, W.; Zheng, Q.-Y. Highly crystalline covalent organic frameworks from flexible building blocks. Chem. Commun. 2016, 52, 4706–4709. [Google Scholar] [CrossRef] [PubMed]
  37. Khalil, S.; Meyer, M.D.; Alazmi, A.; Samani, M.H.; Huang, P.-C.; Barnes, M.; Marciel, A.B.; Verduzco, R. Enabling solution processable COFs through suppression of precipitation during solvothermal synthesis. ACS Nano 2022, 16, 20964–20974. [Google Scholar] [CrossRef] [PubMed]
  38. Xiong, S.; Liu, J.; Wang, Y.; Wang, X.; Chu, J.; Zhang, R.; Gong, M.; Wu, B.J. Solvothermal synthesis of triphenylamine-based covalent organic framework nanofibers with excellent cycle stability for supercapacitor electrodes. Appl. Polym. Sci. 2022, 139, 51510. [Google Scholar] [CrossRef]
  39. 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]
  40. Ren, S.; Bojdys, M.J.; Dawson, R.; Laybourn, A.; Khimyak, Y.Z.; Adams, D.J.; Cooper, A.I. Porous, fluorescent, covalent triazine-based frameworks via room-temperature and microwave-assisted synthesis. Adv. Mater. 2012, 24, 2357–2361. [Google Scholar] [CrossRef]
  41. Wei, H.; Chai, S.; Hu, N.; Yang, Z.; Wei, L.; Wang, L. The microwave-assisted solvothermal synthesis of a crystalline two-dimensional covalent organic framework with high CO2 capacity. Chem. Commun. 2015, 51, 12178–12181. [Google Scholar] [CrossRef]
  42. Kuhn, P.; Antonietti, M.; Thomas, A. Porous, covalent triazine-based frameworks prepared by ionothermal synthesis. Angew. Chem. Int. Ed. 2008, 47, 3450–3453. [Google Scholar] [CrossRef] [PubMed]
  43. Bojdys, M.J.; Jeromenok, J.; Thomas, A.; Antonietti, M. Rational extension of the family of layered, covalent, triazine-based frameworks with regular porosity. Adv. Mater. 2010, 22, 2202–2205. [Google Scholar] [CrossRef] [PubMed]
  44. Guan, X.; Ma, Y.; Li, H.; Yusran, Y.; Xue, M.; Fang, Q.; Yan, Y.; Valtchev, V.; Qiu, S. Fast, ambient temperature and pressure ionothermal synthesis of three-dimensional covalent organic frameworks. J. Am. Chem. Soc. 2018, 140, 4494–4498. [Google Scholar] [CrossRef]
  45. Chandra, S.; Kandambeth, S.; Biswal, B.P.; Lukose, B.; Kunjir, S.M.; Chaudhary, M.; Babarao, R.; Heine, T.; Banerjee, R. Chemically stable multilayered covalent organic nanosheets from covalent organic frameworks via mechanical delamination. J. Am. Chem. Soc. 2013, 135, 17853–17861. [Google Scholar] [CrossRef]
  46. Shinde, D.B.; Aiyappa, H.B.; Bhadra, M.; Biswal, B.P.; Wadge, P.; Kandambeth, S.; Garai, B.; Kundu, T.; Kurungot, S.; Banerjee, R.J. A mechanochemically synthesized covalent organic framework as a proton-conducting solid electrolyte. Mater. Chem. A 2016, 4, 2682–2690. [Google Scholar] [CrossRef]
  47. Dey, K.; Pal, M.; Rout, K.C.; Kunjattu, H.S.; Das, A.; Mukherjee, R.; Kharul, U.K.; Banerjee, R. Selective molecular separation by interfacially crystallized covalent organic framework thin films. J. Am. Chem. Soc. 2017, 139, 13083–13091. [Google Scholar] [CrossRef]
  48. Hao, Q.; Zhao, C.; Sun, B.; Lu, C.; Liu, J.; Liu, M.; Wan, L.-J.; Wang, D. Confined synthesis of two-dimensional covalent organic framework thin films within superspreading water layer. J. Am. Chem. Soc. 2018, 140, 12152–12158. [Google Scholar] [CrossRef]
  49. Zhou, D.; Tan, X.; Wu, H.; Tian, L.; Li, M. Synthesis of C−C Bonded Two-Dimensional Conjugated Covalent Organic Framework Films by Suzuki Polymerization on a Liquid–Liquid Interface. Angew. Chem. Int. Ed. 2019, 58, 1376. [Google Scholar] [CrossRef]
  50. Wan, S.; Guo, J.; Kim, J.; Ihee, H.; Jiang, D. A photoconductive covalent organic framework: Self-condensed arene cubes composed of eclipsed 2D polypyrene sheets for photocurrent generation. Angew. Chem. Int. Ed. 2009, 48, 5439–5442. [Google Scholar] [CrossRef]
  51. Spitler, E.L.; Dichtel, W.R. Lewis acid-catalysed formation of two-dimensional phthalocyanine covalent organic frameworks. Nat. Chem. 2010, 2, 672–677. [Google Scholar] [CrossRef]
  52. Ma, H.; Ren, H.; Meng, S.; Yan, Z.; Zhao, H.; Sun, F.; Zhu, G. A 3D microporous covalent organic framework with exceedingly high C3H8/CH4 and C2 hydrocarbon/CH4 selectivity. Chem. Commun. 2013, 49, 9773–9775. [Google Scholar] [CrossRef] [PubMed]
  53. Dalapati, S.; Jin, E.; Addicoat, M.; Heine, T.; Jiang, D. Highly emissive covalent organic frameworks. J. Am. Chem. Soc. 2016, 138, 5797–5800. [Google Scholar] [CrossRef]
  54. Vitaku, E.; Dichtel, W.R. Synthesis of 2D imine-linked covalent organic frameworks through formal transimination reactions. J. Am. Chem. Soc. 2017, 139, 12911–12914. [Google Scholar] [CrossRef] [PubMed]
  55. Waller, P.J.; AlFaraj, Y.S.; Diercks, C.S.; Jarenwattananon, N.N.; Yaghi, O.M. Conversion of imine to oxazole and thiazole linkages in covalent organic frameworks. J. Am. Chem. Soc. 2018, 140, 9099–9103. [Google Scholar] [CrossRef] [PubMed]
  56. Cusin, L.; Peng, H.; Ciesielski, A.; Samori, P. Chemical Conversion and Locking of the Imine Linkage: Enhancing the Functionality of Covalent Organic Frameworks. Angew. Chem. Int. Ed. 2021, 60, 14236–14250. [Google Scholar] [CrossRef] [PubMed]
  57. Qian, C.; Feng, L.; Teo, W.L.; Liu, J.; Zhou, W.; Wang, D.; Zhao, Y. Imine and imine-derived linkages in two-dimensional covalent organic frameworks. Nat. Rev. Chem. 2022, 6, 881–898. [Google Scholar] [CrossRef]
  58. Zhang, Y.; Farrants, H.; Li, X. Adding a Functional Handle to Nature′ s Building Blocks: The Asymmetric Synthesis of β-Hydroxy-α-Amino Acids. Chem. Asian J. 2014, 9, 1752–1764. [Google Scholar] [CrossRef] [PubMed]
  59. Zhang, Y.; Shen, X.; Feng, X.; Xia, H.; Mu, Y.; Liu, X. Covalent organic frameworks as pH responsive signaling scaffolds. Chem. Commun. 2016, 52, 11088–11091. [Google Scholar] [CrossRef]
  60. Nagai, A.; Chen, X.; Feng, X.; Ding, X.; Guo, Z.; Jiang, D. A squaraine-linked mesoporous covalent organic framework. Angew. Chem. Int. Ed. 2013, 125, 3858–3862. [Google Scholar] [CrossRef]
  61. Ding, N.; Zhou, T.; Weng, W.; Lin, Z.; Liu, S.; Maitarad, P.; Wang, C.; Guo, J. Multivariate Synthetic Strategy for Improving Crystallinity of Zwitterionic Squaraine-Linked Covalent Organic Frameworks with Enhanced Photothermal Performance. Small 2022, 18, 2201275. [Google Scholar] [CrossRef]
  62. Ben, H.; Yan, G.; Liu, H.; Ling, C.; Fan, Y.; Zhang, X. Local spatial polarization induced efficient charge separation of squaraine-linked COF for enhanced photocatalytic performance. Adv. Funct. Mater. 2022, 32, 2104519. [Google Scholar] [CrossRef]
  63. Li, Z.; Zhi, Y.; Feng, X.; Ding, X.; Zou, Y.; Liu, X.; Mu, Y. An azine-linked covalent organic framework: Synthesis, characterization and efficient gas storage. Chem. Eur. J. 2015, 21, 12079–12084. [Google Scholar] [CrossRef] [PubMed]
  64. Alahakoon, S.B.; Thompson, C.M.; Nguyen, A.X.; Occhialini, G.; McCandless, G.T.; Smaldone, R.A. An azine-linked hexaphenylbenzene based covalent organic framework. Chem. Commun. 2016, 52, 2843–2845. [Google Scholar] [CrossRef]
  65. Yu, S.Y.; Mahmood, J.; Noh, H.J.; Seo, J.M.; Jung, S.M.; Shin, S.H.; Im, Y.K.; Jeon, I.Y.; Baek, J.B. Direct synthesis of a covalent triazine-based framework from aromatic amides. Angew. Chem. Int. Ed. 2018, 57, 8438–8442. [Google Scholar] [CrossRef]
  66. Fang, Q.; Zhuang, Z.; Gu, S.; Kaspar, R.B.; Zheng, J.; Wang, J.; Qiu, S.; Yan, Y. Designed synthesis of large-pore crystalline polyimide covalent organic frameworks. Nat. Commun. 2014, 5, 4503. [Google Scholar] [CrossRef]
  67. Fang, Q.; Wang, J.; Gu, S.; Kaspar, R.B.; Zhuang, Z.; Zheng, J.; Guo, H.; Qiu, S.; Yan, Y. 3D porous crystalline polyimide covalent organic frameworks for drug delivery. J. Am. Chem. Soc. 2015, 137, 8352–8355. [Google Scholar] [CrossRef]
  68. Zhuang, X.; Zhao, W.; Zhang, F.; Cao, Y.; Liu, F.; Bi, S.; Feng, X. A two-dimensional conjugated polymer framework with fully sp 2-bonded carbon skeleton. Polym. Chem. 2016, 7, 4176–4181. [Google Scholar] [CrossRef]
  69. Jin, E.; Asada, M.; Xu, Q.; Dalapati, S.; Addicoat, M.A.; Brady, M.A.; Xu, H.; Nakamura, T.; Heine, T.; Chen, Q. Two-dimensional sp2 carbon–conjugated covalent organic frameworks. Science 2017, 357, 673–676. [Google Scholar] [CrossRef] [PubMed]
  70. Zhao, Y.; Liu, H.; Wu, C.; Zhang, Z.; Pan, Q.; Hu, F.; Wang, R.; Li, P.; Huang, X.; Li, Z. Fully conjugated two-dimensional sp2-carbon covalent organic frameworks as artificial photosystem I with high efficiency. Angew. Chem. Int. Ed. 2019, 58, 5376–5381. [Google Scholar] [CrossRef] [PubMed]
  71. Lyu, H.; Diercks, C.S.; Zhu, C.; Yaghi, O.M. Porous crystalline olefin-linked covalent organic frameworks. J. Am. Chem. Soc. 2019, 141, 6848–6852. [Google Scholar] [CrossRef] [PubMed]
  72. Zhang, B.; Wei, M.; Mao, H.; Pei, X.; Alshmimri, S.A.; Reimer, J.A.; Yaghi, O.M. Crystalline dioxin-linked covalent organic frameworks from irreversible reactions. J. Am. Chem. Soc. 2018, 140, 12715–12719. [Google Scholar] [CrossRef]
  73. Guan, X.; Li, H.; Ma, Y.; Xue, M.; Fang, Q.; Yan, Y.; Valtchev, V.; Qiu, S. Chemically stable polyarylether-based covalent organic frameworks. Nat. Chem. 2019, 11, 587–594. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, C.; Zhang, S.; Yan, Y.; Xia, F.; Huang, A.; Xian, Y. Highly fluorescent polyimide covalent organic nanosheets as sensing probes for the detection of 2, 4, 6-trinitrophenol. ACS Appl. Mater. Interfaces 2017, 9, 13415–13421. [Google Scholar] [CrossRef] [PubMed]
  75. Das, G.; Garai, B.; Prakasam, T.; Benyettou, F.; Varghese, S.; Sharma, S.K.; Gándara, F.; Pasricha, R.; Baias, M.; Jagannathan, R. Fluorescence turn on amine detection in a cationic covalent organic framework. Nat. Commun. 2022, 13, 3904. [Google Scholar] [CrossRef] [PubMed]
  76. Zhang, J.-L.; Yao, L.-Y.; Yang, Y.; Liang, W.-B.; Yuan, R.; Xiao, D.-R. Conductive covalent organic frameworks with conductivity-and pre-reduction-enhanced electrochemiluminescence for ultrasensitive biosensor construction. Anal. Chem. 2022, 94, 3685–3692. [Google Scholar] [CrossRef] [PubMed]
  77. Kurisingal, J.F.; Kim, H.; Choe, J.H.; Hong, C.S. Covalent organic framework-based catalysts for efficient CO2 utilization reactions. Coord. Coord. Chem. Rev. 2022, 473, 214835. [Google Scholar] [CrossRef]
  78. Bai, Y.; Liu, Y.; Liu, M.; Wang, X.; Shang, S.; Gao, W.; Du, C.; Qiao, Y.; Chen, J.; Dong, J. Near-Equilibrium Growth of Chemically Stable Covalent Organic Framework/Graphene Oxide Hybrid Materials for the Hydrogen Evolution Reaction. Angew. Chem. Int. Ed. 2022, 61, e202113067. [Google Scholar] [CrossRef]
  79. Liu, S.; Wang, M.; He, Y.; Cheng, Q.; Qian, T.; Yan, C. Covalent organic frameworks towards photocatalytic applications: Design principles, achievements, and opportunities. Coord. Chem. Rev. 2023, 475, 214882. [Google Scholar] [CrossRef]
  80. Yu, A.; Pan, Q.; Zhang, M.; Xie, D.; Tang, Y. Fast rate and long life potassium-ion based dual-ion battery through 3D porous organic negative electrode. Adv. Funct. Mater. 2020, 30, 2001440. [Google Scholar] [CrossRef]
  81. Yang, X.; Hu, Y.; Dunlap, N.; Wang, X.; Huang, S.; Su, Z.; Sharma, S.; Jin, Y.; Huang, F.; Wang, X. A truxenone-based covalent organic framework as an all-solid-state lithium-ion battery cathode with high capacity. Angew. Chem. Int. Ed. 2020, 59, 20385–20389. [Google Scholar] [CrossRef]
  82. Haldar, S.; Rase, D.; Shekhar, P.; Jain, C.; Vinod, C.P.; Zhang, E.; Shupletsov, L.; Kaskel, S.; Vaidhyanathan, R. Incorporating Conducting Polypyrrole into a Polyimide COF for Carbon-Free Ultra-High Energy Supercapacitor. Adv. Energy Mater. 2022, 12, 2200754. [Google Scholar] [CrossRef]
  83. Shah, R.; Ali, S.; Raziq, F.; Ali, S.; Ismail, P.M.; Shah, S.; Iqbal, R.; Wu, X.; He, W.; Zu, X.; et al. Exploration of metal organic frameworks and covalent organic frameworks for energy-related applications. Coord. Chem. Rev. 2023, 477, 214968. [Google Scholar] [CrossRef]
  84. Ma, L.; Wang, S.; Feng, X.; Wang, B. Recent advances of covalent organic frameworks in electronic and optical applications. Chin. Chem. Lett. 2016, 27, 1383–1394. [Google Scholar] [CrossRef]
  85. Sun, B.; Li, X.; Feng, T.; Cai, S.; Chen, T.; Zhu, C.; Zhang, J.; Wang, D.; Liu, Y. Resistive switching memory performance of two-dimensional polyimide covalent organic framework films. ACS Appl. Mater. Interfaces 2020, 12, 51837–51845. [Google Scholar] [CrossRef] [PubMed]
  86. Hao, Q.; Li, Z.J.; Bai, B.; Zhang, X.; Zhong, Y.W.; Wan, L.J.; Wang, D. A covalent organic framework film for three-state near-infrared electrochromism and a molecular logic gate. Angew. Chem. Int. Ed. 2021, 133, 12606–12611. [Google Scholar] [CrossRef]
  87. Zhou, K.; Jia, Z.; Zhou, Y.; Ding, G.; Ma, X.-Q.; Niu, W.; Han, S.-T.; Zhao, J.; Zhou, Y.J. Covalent Organic Frameworks for Neuromorphic Devices. Phys. Chem. Lett. 2023, 14, 7173–7192. [Google Scholar] [CrossRef]
  88. Yu, H.; Zhou, P.K.; Chen, X. Intramolecular Hydrogen Bonding Interactions Induced Enhancement in Resistive Switching Memory Performance for Covalent Organic Framework-Based Memristors. Adv. Funct. Mater. 2023, 33, 2308336. [Google Scholar] [CrossRef]
  89. Ding, G.; Zhao, J.; Zhou, K.; Zheng, Q.; Han, S.-T.; Peng, X.; Zhou, Y. Porous crystalline materials for memories and neuromorphic computing systems. Chem. Soc. Rev. 2023, 52, 7071–7136. [Google Scholar] [CrossRef] [PubMed]
  90. Zhou, P.K.; Yu, H.; Huang, W.; Chee, M.Y.; Wu, S.; Zeng, T.; Lim, G.J.; Xu, H.; Yu, Z.; Li, H. Photoelectric Multilevel Memory Device based on Covalent Organic Polymer Film with Keto–Enol Tautomerism for Harsh Environments Applications. Adv. Funct. Mater. 2023, 2306593. [Google Scholar] [CrossRef]
  91. Zhou, P.K.; Yu, H.; Li, Y.; Yu, H.; Chen, Q.; Chen, X. Recent advances in covalent organic polymers-based thin films as memory devices. J. Polym. Sci. 2023, 1. [Google Scholar] [CrossRef]
  92. Gu, Q.; Zha, J.; Chen, C.; Wang, X.; Yao, W.; Liu, J.; Kang, F.; Yang, J.; Li, Y.Y.; Lei, D. Constructing Chiral Covalent-Organic Frameworks for Circularly Polarized Light Detection. Adv. Mater. 2023, 2306414. [Google Scholar] [CrossRef]
  93. Xu, F.; Xu, H.; Chen, X.; Wu, D.; Wu, Y.; Liu, H.; Gu, C.; Fu, R.; Jiang, D. Radical covalent organic frameworks: A general strategy to immobilize open-accessible polyradicals for high-performance capacitive energy storage. Angew. Chem. Int. Ed. 2015, 127, 6918–6922. [Google Scholar] [CrossRef]
  94. Xu, F.; Yang, S.; Chen, X.; Liu, Q.; Li, H.; Wang, H.; Wei, B.; Jiang, D. Energy-storage covalent organic frameworks: Improving performance via engineering polysulfide chains on walls. Chem. Sci. 2019, 10, 6001–6006. [Google Scholar] [CrossRef] [PubMed]
  95. Wu, J.; Xu, F.; Li, S.; Ma, P.; Zhang, X.; Liu, Q.; Fu, R.; Wu, D. Porous polymers as multifunctional material platforms toward task-specific applications. Adv. Mater. 2019, 31, 1802922. [Google Scholar] [CrossRef] [PubMed]
  96. Zhuang, R.; Zhang, X.; Qu, C.; Xu, X.; Yang, J.; Ye, Q.; Liu, Z.; Kaskel, S.; Xu, F.; Wang, H. Fluorinated porous frameworks enable robust anode-less sodium metal batteries. Sci. Adv. 2023, 9, eadh8060. [Google Scholar] [CrossRef] [PubMed]
  97. Wu, T.; He, Q.; Liu, Z.; Shao, B.; Liang, Q.; Pan, Y.; Huang, J.; Peng, Z.; Liu, Y.; Zhao, C.J. Tube wall delamination engineering induces photogenerated carrier separation to achieve photocatalytic performance improvement of tubular g-C3N4. Hazard. Mater. 2022, 424, 127177. [Google Scholar] [CrossRef] [PubMed]
  98. Gu, C.-C.; Xu, F.-H.; Zhu, W.-K.; Wu, R.-J.; Deng, L.; Zou, J.; Weng, B.-C.; Zhu, R.-L. Recent advances on covalent organic frameworks (COFs) as photocatalysts: Different strategies for enhancing hydrogen generation. Chem. Commun. 2023, 59, 7302–7320. [Google Scholar] [CrossRef] [PubMed]
  99. Reza, M.S.; Ahmad, N.B.H.; Afroze, S.; Taweekun, J.; Sharifpur, M.; Azad, A.K. Hydrogen Production from Water Splitting through Photocatalytic Activity of Carbon-Based Materials. Chem. Eng. Technol. 2023, 46, 420–434. [Google Scholar] [CrossRef]
  100. Hou, H.; Zeng, X.; Zhang, X. Production of Hydrogen Peroxide by Photocatalytic Processes. Angew. Chem. Int. Ed. 2020, 59, 17356–17376. [Google Scholar] [CrossRef]
  101. Di, T.; Xu, Q.; Ho, W.; Tang, H.; Xiang, Q.; Yu, J. Review on metal sulphide-based Z-scheme photocatalysts. ChemCatChem 2019, 11, 1394–1411. [Google Scholar] [CrossRef]
  102. Teranishi, M.; Naya, S.-I.; Tada, H. In situ liquid phase synthesis of hydrogen peroxide from molecular oxygen using gold nanoparticle-loaded titanium (IV) dioxide photocatalyst. J. Am. Chem. Soc. 2010, 132, 7850–7851. [Google Scholar] [CrossRef] [PubMed]
  103. Tsukamoto, D.; Shiro, A.; Shiraishi, Y.; Sugano, Y.; Ichikawa, S.; Tanaka, S.; Hirai, T. Photocatalytic H2O2 Production from Ethanol/O2 System Using TiO2 Loaded with Au–Ag Bimetallic Alloy Nanoparticles. ACS Catal. 2012, 2, 599–603. [Google Scholar] [CrossRef]
  104. 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] [PubMed]
  105. Zhou, S.; Hu, H.; Hu, H.; Jiang, Q.; Xie, H.; Li, C.; Gao, S.; Kong, Y.; Hu, Y. Unveiling the latent reactivity of imines on pyridine-functionalized covalent organic frameworks for H2O2 photosynthesis. Sci. China Mater. 2023, 66, 1837–1846. [Google Scholar] [CrossRef]
  106. Kou, M.; Wang, Y.; Xu, Y.; Ye, L.; Huang, Y.; Jia, B.; Li, H.; Ren, J.; Deng, Y.; Chen, J.; et al. Molecularly Engineered Covalent Organic Frameworks for Hydrogen Peroxide Photosynthesis. Angew. Chem. Int. Ed. 2022, 61, e202200413. [Google Scholar] [CrossRef] [PubMed]
  107. Wang, H.; Yang, C.; Chen, F.; Zheng, G.; Han, Q. A Crystalline Partially Fluorinated Triazine Covalent Organic Framework for Efficient Photosynthesis of Hydrogen Peroxide. Angew. Chem. Int. Ed. 2022, 61, e202202328. [Google Scholar] [CrossRef] [PubMed]
  108. Mou, Y.; Wu, X.; Qin, C.; Chen, J.; Zhao, Y.; Jiang, L.; Zhang, C.; Yuan, X.; Huixiang Ang, E.; Wang, H. Linkage Microenvironment of Azoles-Related Covalent Organic Frameworks Precisely Regulates Photocatalytic Generation of Hydrogen Peroxide. Angew. Chem. Int. Ed. 2023, 62, e202309480. [Google Scholar] [CrossRef] [PubMed]
  109. 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]
  110. 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]
  111. 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. 2020, 32, e1904433. [Google Scholar] [CrossRef]
  112. 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. 2022, 34, e2107480. [Google Scholar] [CrossRef] [PubMed]
  113. 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] [PubMed]
  114. Wu, C.; Teng, Z.; Yang, C.; Chen, F.; Yang, H.B.; Wang, L.; Xu, H.; Liu, B.; Zheng, G.; Han, Q. Polarization Engineering of Covalent Triazine Frameworks for Highly Efficient Photosynthesis of Hydrogen Peroxide from Molecular Oxygen and Water. Adv. Mater. 2022, 34, e2110266. [Google Scholar] [CrossRef] [PubMed]
  115. 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]
Figure 1. The advantages of COFs for the photosynthesis of H2O2.
Figure 1. The advantages of COFs for the photosynthesis of H2O2.
Polymers 16 00659 g001
Figure 2. (a) Schematic illustration and (b) Corresponding energy diagrams of the oxygen reduction and water oxidation involved in H2O2 photosynthesis. Reproduced with permission [1].
Figure 2. (a) Schematic illustration and (b) Corresponding energy diagrams of the oxygen reduction and water oxidation involved in H2O2 photosynthesis. Reproduced with permission [1].
Polymers 16 00659 g002
Figure 3. (a) Synthetic process of sulfone-modified COFs. (b) Photocatalytic H2O2 production rate of C-COFs (black), S-COFs (red), and FS-COFs (blue). (c) UV/Vis spectrum of FS-COFs and AQY for H2O2 production. Reproduced with permission [113]. Copyright 2023, Wiley-VCH.
Figure 3. (a) Synthetic process of sulfone-modified COFs. (b) Photocatalytic H2O2 production rate of C-COFs (black), S-COFs (red), and FS-COFs (blue). (c) UV/Vis spectrum of FS-COFs and AQY for H2O2 production. Reproduced with permission [113]. Copyright 2023, Wiley-VCH.
Polymers 16 00659 g003
Figure 4. (a) Synthetic route of thiourea-functionalized COFs. (b) Relative OH bending intensity of thiourea-functionalized COFs. (c) H2O2 generation rate. (d) Weight differential index of fragments in excited states 4–9 of Bpt-CTF. (e) Schematic diagram of the catalytic pathway. Reproduced with permission [114]. Copyright 2022, Wiley-VCH.
Figure 4. (a) Synthetic route of thiourea-functionalized COFs. (b) Relative OH bending intensity of thiourea-functionalized COFs. (c) H2O2 generation rate. (d) Weight differential index of fragments in excited states 4–9 of Bpt-CTF. (e) Schematic diagram of the catalytic pathway. Reproduced with permission [114]. Copyright 2022, Wiley-VCH.
Polymers 16 00659 g004
Figure 5. (a) Schematic diagram of the synthesis process. (b) H2O2 production yield. (c) H2O2 production rate. (d) AQY of TpDz at chosen wavelengths: 420 (purple), 450 (light blue), 500 (green), and 600 (red) nm. Reproduced with permission. Copyright 2023 [104], Wiley-VCH.
Figure 5. (a) Schematic diagram of the synthesis process. (b) H2O2 production yield. (c) H2O2 production rate. (d) AQY of TpDz at chosen wavelengths: 420 (purple), 450 (light blue), 500 (green), and 600 (red) nm. Reproduced with permission. Copyright 2023 [104], Wiley-VCH.
Polymers 16 00659 g005
Figure 6. (a) Synthetic process of fluorinated COFs. (b) N2 sorption isotherms of obtained COFs. (c) Specific surface areas and pore volumes in H-COF, TF-COF, and TF50-COF Using BET and Langmuir Models. (d) Crystal stacking energies of TF-COF and H-COF. (e) Photocatalytic H2O2 production rate for 1 h. (f) Photocatalytic H2O2 production rate for 5 h. (g) The SCC efficiencies of synthesized COFs. Reproduced with permission [107]. Copyright 2022, Wiley-VCH.
Figure 6. (a) Synthetic process of fluorinated COFs. (b) N2 sorption isotherms of obtained COFs. (c) Specific surface areas and pore volumes in H-COF, TF-COF, and TF50-COF Using BET and Langmuir Models. (d) Crystal stacking energies of TF-COF and H-COF. (e) Photocatalytic H2O2 production rate for 1 h. (f) Photocatalytic H2O2 production rate for 5 h. (g) The SCC efficiencies of synthesized COFs. Reproduced with permission [107]. Copyright 2022, Wiley-VCH.
Polymers 16 00659 g006
Figure 7. (a) Synthetic route of BTEA-COF (red) and EBA-COF (blue). (b) UV-DRS of EBA-COF (blue) and BTEA-COF (red). (c) Mott-Schottky plot of the BTEA-COF. (d) Reactions of EBA-COF under varied gas atmospheres. (e) Photocatalytic production of H2O2 using EBA-COF in the presence of KI, CuSO4, and p-benzoquinone. Reproduced with permission [115]. Copyright 2022, American Chemical Society.
Figure 7. (a) Synthetic route of BTEA-COF (red) and EBA-COF (blue). (b) UV-DRS of EBA-COF (blue) and BTEA-COF (red). (c) Mott-Schottky plot of the BTEA-COF. (d) Reactions of EBA-COF under varied gas atmospheres. (e) Photocatalytic production of H2O2 using EBA-COF in the presence of KI, CuSO4, and p-benzoquinone. Reproduced with permission [115]. Copyright 2022, American Chemical Society.
Polymers 16 00659 g007
Table 1. Representative COF photocatalysts for H2O2 generation (SCC, AQY, EtOH, BA, and MeOH are abbreviations for solar-to-chemical energy conversion, apparent quantum yield, ethanol, benzyl alcohol, and methanol, respectively; “/” means the data is not provided in the literature).
Table 1. Representative COF photocatalysts for H2O2 generation (SCC, AQY, EtOH, BA, and MeOH are abbreviations for solar-to-chemical energy conversion, apparent quantum yield, ethanol, benzyl alcohol, and methanol, respectively; “/” means the data is not provided in the literature).
EntryLight Source (nm)H2O2 Generation Rate (μmol·h−1·g−1)SCC Efficiency (%)AQY (%) at 420 nmReaction PathwaySolventRef.
TAPD-omeCOF420–70097 ± 10//2e ORRH2O:EtOH (9:1)[20]
TAPD-meCOF420–70097 ± 10//2e ORRH2O:EtOH (9:1)[20]
TF50-COFλ > 40017390.175.1Two-step
1e ORR
H2O:EtOH (9:1)[21]
TTF-BT-COF420–700 27600.4911.192e ORR
2e WOR
H2O[22]
EBA-COFλ = 420 1820
2550
/4.4Two-step
1e ORR
H2O:EtOH (9:1)
H2O:BA (9:1)
[23]
HEP-TAPT-COFλ > 4201750 0.65 15.35 2e ORRH2O[24]
Cof-TpDzλ > 42073270.6211.92e ORRH2O[25]
Py-Da-COF420–700461
682
1242
/
2.4
4.5
2e ORRH2O
H2O:EtOH (9:1)
H2O:BA (9:1)
[14]
FS-COFλ > 4203904.2/6.212e ORRH2O[26]
TZ-COFλ > 420268
386
4951
0.0360.62e ORRH2O
H2O:MeOH (1:1)
H2O:BA (1:1)
[27]
TAPT-TFPA COFsXe lamp21430.826.5 at 400 nmTwo-step
1e ORR
H2O:EtOH (9:1)[28]
CHF-DPDAλ > 4201725 0.78162e ORR
2e WOR
H2O[29]
CTF-BDDBNλ > 42097.20.14/2e ORR
2e WOR
H2O:MeOH (9:1)[30]
COF-TfpBpyλ > 420 6950.578.1 2e ORR
2e WOR
H2O[31]
COF-N32λ > 4206050.316.2 at 459 nm2e ORR
2e WOR
H2O[32]
Bpt-CTF350–7803268.1/6.4 2e ORRH2O[33]
sonoCOF-F2λ > 4202736/4.8Two-step
1e ORR
H2O[34]
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

Tan, D.; Fan, X. COF-Based Photocatalysts for Enhanced Synthesis of Hydrogen Peroxide. Polymers 2024, 16, 659. https://doi.org/10.3390/polym16050659

AMA Style

Tan D, Fan X. COF-Based Photocatalysts for Enhanced Synthesis of Hydrogen Peroxide. Polymers. 2024; 16(5):659. https://doi.org/10.3390/polym16050659

Chicago/Turabian Style

Tan, Deming, and Xuelin Fan. 2024. "COF-Based Photocatalysts for Enhanced Synthesis of Hydrogen Peroxide" Polymers 16, no. 5: 659. https://doi.org/10.3390/polym16050659

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