Open Access Article
Qianshuo Nan†
ab,
Jing Ning†
b,
Bing Han
c,
Hongtao Weib,
Xuefeng Wang
*b,
Ying-Ying Gu*a,
Shengxiang Zhoud,
Guangqiang Caob,
Guangze Zhangb,
Xuehui Lib,
Yonggang Jia*e and
Long Hao
*b
aCollege of Chemistry and Chemical Engineering, China University of Petroleum (East China), Qingdao 266580, P. R. China. E-mail: yingyinggu@upc.edu.cn
bCollege of Chemistry and Pharmaceutical Sciences, Qingdao Agricultural University, No. 700 Changcheng Road, Qingdao 266109, P. R. China. E-mail: wxf@qau.edu.cnhaol@qau.edu.cn
cMOE Key Laboratory of Resources and Environmental Systems Optimization, College of Environmental Science and Engineering, North China Electric Power University, Beijing, 102206, P.R. China
dMarine Science Research Institute of Shandong Province, 7 Youyun Road, Qingdao, 266104, China
eShandong Provincial Key Laboratory of Marine Environment and Geological Engineering, Key Laboratory of Marine Environment and Ecology, Ocean University of China, Qingdao 266100, China. E-mail: yonggang@ouc.edu.cn
First published on 8th November 2025
Simultaneously improving charge carrier separation and surface reaction efficiency is crucial for enhancing the photocatalytic H2O2 production efficiency of covalent organic frameworks (COFs). Here, a heteroatom-lock strategy is introduced into the acceptor structure of COFs, with the “lock” effect to enhance the coplanarity and conjugation, and the “heteroatom” effect to improve the O2 adsorption. It turns out that the photocatalytic H2O2 production yield of the N-heteroatom locked COF (2.08 mmol g−1 h−1 under pure water and air conditions) is 2.1 times that of the S-heteroatom locked COF and 4.7 times that of the original COF. Experimental results and theoretical calculations reveal that the heteroatom-lock-induced H2O2 production enhancement of COFs is attributed to their lower exciton binding energy (Eb) and smaller charge transfer resistance, together with the bigger O2 adsorption energy and lower transition state energy of the intermediates. Additionally, a novel gas diffusion reaction system is developed to further improve the O2 diffusion efficiency, which not only enhances the photocatalytic H2O2 production yield to 4.06 mmol g−1 h−1, but also realizes the immobilization and efficient recycling of the COF catalyst. This study provides new insights into the rational design of COF-based photocatalysts, and offers a novel approach for the reaction system of photocatalytic H2O2 production.
Previous studies have shown that traditional inorganic photocatalysts, such as TiO2,11–13 BiVO4,14,15 and metal sulfides,16,17 exhibit limited performance in photocatalytic H2O2 production. In contrast, organic polymeric photocatalysts, including graphitic carbon nitride (C3N4), conjugated organic polymers (COPs), and covalent organic frameworks (COFs) display significantly superior catalytic activity and stability.18–20 Notably, COF-based catalysts have received extensive attention since their first reported application in photocatalytic H2O2 production in 2020.21 This is not only attributed to the high specific surface area and ordered through-pore structure of COFs, which provide a large number of accessible active sites, but also to their molecular-level controllable framework structures and linkage types, which enable precise modulation of the catalyst's photoresponse, as well as the thermodynamics and kinetics of the reaction processes.22–24
At present, the photocatalytic H2O2 production efficiency of COFs still presents substantial room for improvement, which essentially needs in-depth investigation of the relationship between the COF structure and the fundamental photocatalytic processes (photoabsorption, carrier separation/transport, and surface redox reactions).25 Normally, photoabsorption can be properly tuned by the monomer and linkage types of the COF.26–31 However, the carrier separation/transport is more complicated, which is not only decided by the skeleton structure of the COF,32–37 but also affected by its microstructures, including conjugation,38–40 heteroatom-containing species (e.g. S/O/Se/N),41,42 position/content of nitrogen,43–45 or surface functional groups46–49 in the COF monomers. As for the surface redox reactions, the mass transport requires greater attention,50–52 especially for the O2-related species, in which heteroatoms with high oxygen affinity (e.g. N/S) are usually introduced,53–56 and a flow reaction system has also been developed recently.57 Therefore, to improve the overall efficiency of COF catalysts, it is necessary to consider the synergistic effect of the monomer structure on different photocatalytic processes. Meanwhile, developing a more reasonable reaction system may also be an effective approach.
Herein, a heteroatom-lock strategy is applied in the acceptor part of the COF (Scheme 1a), specifically, S (for 3,7-diaminodibenzothiophene, DBT) or N (for 6-phenylphenanthridine-3,8-diamine, PhPD) atom is used to lock the rotation of the diaminobiphenyl (BPh) acceptor (the corresponding donor part of the COFs is benzotrithiophenetricarbaldehyde, BTT, a widely-used donor developed by our group58–61), in which the improvement of coplanarity (“lock” effect) aims to enhance the charge carrier separation/transport; meanwhile, the introduction of the S/N heteroatom is to improve the O2 adsorption capacity (see the simulated O2 adsorption energy of the structural units of COFs in Scheme 1b: −0.13 eV for BTT-BPh, −0.16 eV for S-locked BTT-DBT, and −0.21 eV for N-locked BTT-PhPD). It turns out that the photocatalytic H2O2 production yield of the N-locked COF (2.08 mmol g−1 h−1 under pure water and air conditions) is 2.1 times that of the S-locked COF and 4.7 times that of the original COF. Experimental results and theoretical calculations reveal that the heteroatom-lock-induced H2O2 production enhancement of the COF is attributed to its lower exciton binding energy (Eb) and smaller charge transfer resistance, together with the bigger O2 adsorption energy and lower transition state energy of the intermediates. Additionally, a gas diffusion system (a simplified system similar to the O2-involved heterogeneous electrocatalytic system62) is also introduced to improve the O2 diffusion efficiency, which not only further enhances the photocatalytic H2O2 yield to 4.06 mmol g−1 h−1, but also realizes the immobilization and efficient recycling of the COF catalyst. This work elucidates the synergistic effect of the heteroatom-locked acceptor and the gas diffusion reaction system toward improvement of the fundamental photocatalytic processes, providing valuable insights for the rational design of efficient COF-based photocatalysts and reaction systems.
Surface area and porosity of the COFs were evaluated through N2 adsorption–desorption isotherms at 77 K (Fig. 1g–i). All the COFs demonstrate similar and large Brunauer–Emmett–Teller (BET) surface areas: 1564 m2 g−1 for BTT-BPh, 1492 m2 g−1 for BTT-DBT, and 1538 m2 g−1 for BTT-PhPD (Fig. S2), and the corresponding average pore sizes are 3.2 nm (BTT-BPh), 3.02 nm (BTT-DBT), and 2.59 nm (BTT-PhPD), respectively (insets in Fig. 1g–i). The high surface areas and abundant mesopores provide numerous active sites for the photocatalytic reactions, while the similar surface areas enable a direct comparison of the structure–function relationship among the three COFs. FT-IR spectra of the COFs in Fig. 1j and S3 exhibit the reduction of amino group (–NH2) stretching vibration peaks at 3200–3400 cm−1 and carbonyl group (–C
O) stretching vibration peaks at ∼1660 cm−1 compared to the monomers, while displaying the appearance of imine (–C
N–) stretching vibration peaks at 1605–1617 cm−1, which confirms the formation of imine bonds through the condensation of the aldehyde and amine monomers. Besides, the thermogravimetric analysis (TGA) data demonstrate that all the COFs exhibit no obvious weight-loss before 230 °C, displaying their good thermal stability (Fig. S4). Water contact angle measurements for the three COFs indicate that the incorporation of heteroatoms leads to a moderate enhancement in the overall hydrophilicity of the materials (Fig. S5).
The light-harvesting properties of the COFs were analysed using UV-vis diffuse reflectance spectroscopy (UV-vis DRS) (Fig. 1k). The COFs all exhibit strong light absorption in the visible-light region (digital photos in Fig. S6), and with the heteroatom-lock, the light absorption edges of BTT-DBT and BTT-PhPD all red-shift. Correspondingly, the bandgaps of BTT-PhPD (2.02 eV) and BTT-DBT (2.05 eV) are smaller than that of BTT-BPh (2.13 eV) (see Tauc plots as the inset in Fig. 1k). Comparative analysis of the N 1s XPS spectra for the three COFs (Fig. S7) reveals that the imine N binding energies in both the N-locked BTT-PhPD (398.61 eV) and S-locked BTT-DBT (398.52 eV) are lower than that of pristine BTT-BPh (398.72 eV), which indicates a higher electron density around the N atoms. The enhanced electron density at the imine N can be rationally explained by the extended conjugation within the heteroatom-locked skeleton, which facilitates more efficient electron delocalization onto the N atoms. The red-shifted absorption edges and the decreased imine N binding energies indirectly suggest an extension of the π-conjugation. The MS curves of the COFs all exhibit a positive slope, indicating their n-type semiconductor nature (Fig. S8). The MS curves also indicate that the conduction band minimum (CBM) potentials of BTT-BPh, BTT-DBT, and BTT-PhPD are −0.88 V, −0.67 V and −0.65 V, respectively. According to the equation Eg = EVBM − ECBM, the valence band maximum (VBM) edge potentials are calculated to be 1.25 V, 1.38 V and 1.37 V, respectively. Consequently, the band structure diagrams of the COFs were constructed, which reveal that all the COFs can thermodynamically drive photocatalytic H2O2 production.
The above structural characterization studies of the COFs demonstrate their highly ordered structures, large surface areas, abundant mesopores, and proper band structures, which are all suitable for photocatalytic H2O2 synthesis. Subsequently, their photocatalytic performances were systematically evaluated under simulated sunlight irradiation (λ > 420 nm) in pure water (without sacrificial agents) and air. As shown in Fig. 2a, compared with BTT-BPh (0.44 mmol g−1 h−1), the heteroatom-locked COFs, BTT-DBT (0.98 mmol g−1 h−1) and BTT-PhPD (2.08 mmol g−1 h−1), show much higher H2O2 production rates (2.1 and 4.7 times, respectively), which is comparable to that of most reported COF systems (Table S4). The corresponding apparent quantum efficiency (AQY) was also evaluated at various wavelengths from 400 to 600 nm (Fig. 2b),63 which shows that the COFs all possess the highest AQY at 420 nm, and BTT-PhPD has the highest AQY of 4.0% compared with BTT-DBT (2.6%) and BTT-BPh (1.9%). The solar-to-chemical conversion (SCC) efficiencies of the COFs are 0.12% (BTT-BPh), 0.16% (BTT-DBT) and 0.32% (BTT-PhPD), respectively. Besides, the photocatalytic performance of the COFs was also tested under an O2 atmosphere instead of air, which only displayed a little improvement in the H2O2 yield (typically from 2.08 to 2.42 mmol g−1 h−1 for BTT-PhPD, Fig. S9, summarized in Fig. 2c). When benzyl alcohol was used as the hole scavenger,64,65 the H2O2 yield of all the COFs increased nearly threefold (typically 2.08 to 6.65 mmol g−1 h−1 for BTT-PhPD, Fig. S10 and 2c). It can be seen that the H2O2 production rate consistently followed the order BTT-PhPD > BTT-DBT > BTT-BPh, which demonstrates the advantages of COFs with a heteroatom-locked acceptor in photocatalytic H2O2 production, especially when using N as the locking atom. Additionally, after the photocatalytic test, all the COFs still retain a certain degree of crystallinity (Fig. S11), and similar FT-IR spectra (Fig. S12), confirming their great structural stability. The H2O2 degradation experiments also indicated that the three COFs exhibited almost no degradation of H2O2 (Fig. S13). These all ensure a sustained COF-based photocatalytic H2O2 production.
Normally, photocatalytic H2O2 production may involve the oxygen reduction reaction (ORR) and water oxidation reaction (WOR), each of which also has several pathways.66–68 In order to investigate the specific reaction processes of these COF-based photocatalysts H2O2 production, a series of capture experiments, in situ characterization experiments, and electrochemical experiments were conducted. Under an argon (Ar) atmosphere, all the COFs produced negligible amounts of H2O2 (Fig. 2c), indicating that O2 is the key factor controlling H2O2 generation. When p-benzoquinone (p-BQ, 10 mM) was added as the superoxide radical (˙O2−) scavenger in pure water and air, a sharp decrease in H2O2 production rates was observed for all three COFs (typically from 2.08 to 0.05 mmol g−1 h−1 for BTT-PhPD, Fig. 2c). Additionally, in the nitroblue tetrazolium (NBT) experiment, the NBT content in the solution decreased to varying extents with prolonged illumination (Fig. S14), further confirming the generation of ˙O2− during the reaction process.69 Furthermore, within the same time frame, the NBT content in the solution with BTT-PhPD decreased the most, demonstrating that BTT-PhPD generates more ˙O2− than BTT-DBT and BTT-BPh, which could be the key to its high H2O2 yield. The production of ˙O2− was also demonstrated by electron paramagnetic resonance (EPR) experiments, where COFs of the same mass exhibited clear DMPO–˙O2− characteristic signals upon the addition of a spin trap (DMPO) under light irradiation, with BTT-PhPD showing the highest intensity (Fig. 2d). Diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) of BTT-PhPD in H2O vapor and O2 atmosphere further revealed the ORR reaction intermediates: the intensities of the O–O characteristic peak at 895 cm−1, the ˙O2−/˙OOH characteristic peak at 1155 cm−1, and the *HOOH characteristic peak at 1356 cm−1 all increase over time under illumination (Fig. 2e), further confirming the indirect ORR pathway for H2O2 production.70,71 Silver nitrate (AgNO3) was also used as the electron sacrificial agent to investigate the WOR half-reaction, but almost no H2O2 was detected in the AgNO3 solution (10 mM) under Ar atmosphere (Fig. 2c), suggesting that the WOR process contributes minimally to H2O2 production, and indicating that only the oxidation of H2O to O2 occurs in the WOR process. Furthermore, the reaction pathway for the 2e− ORR can also be determined through electrochemical experiments on a rotating disk electrode (RDE). The linear sweep voltammetry (LSV) curves were obtained at various disk rotation speeds ranging from 400 to 2000 rpm (Fig. S15). Using the Koutecky–Levich method, the electron transfer numbers for the COFs were calculated to be 1.73 (BTT-BPh), 1.82 (BTT-DBT), and 2.08 (BTT-PhPD) (Fig. 2f).72,73 Based on the above results, it is inferred that the generation of H2O2 in the three COFs primarily involves indirect 2 e− ORR (O2 + e− → ˙O2−; ˙O2− + e− + 2H+ → H2O2) and 4 e− WOR (2H2O − 4 e− → O2 + 4H+) processes (typical schematic in Fig. S16).74,75
To further explore the reasons for the heteroatom-lock effect on the promotion of photocatalytic H2O2 production, the differences of the fundamental photocatalytic processes between the three COFs were characterized systematically. First, temperature-dependent photoluminescence (PL) experiments were performed to investigate the charge carrier dynamics of the COFs. The emission intensity of all COFs decreases with increasing temperature, due to the progression of the thermally activated nonradiative recombination process under resonant excitation (Fig. S17).76 The calculated exciton binding energy (Eb) results indicate that BTT-PhPD (164.2 meV) has a much lower Eb compared to BTT-DBT (204.6 meV) and BTT-BPh (235.8 meV) (Fig. 3a), suggesting that excitons in BTT-PhPD are more readily dissociated, facilitating their participation in photocatalytic reactions.77 Furthermore, time-resolved PL analysis in the nanosecond to microsecond time domain reveals that the fluorescence lifetimes of BTT-PhPD (1.89 ns) and BTT-DBT (1.76 ns) are nearly double that of BTT-BPh (1.07 ns) (Fig. 3b), suggesting a greater suppression of photogenerated carrier recombination in BTT-PhPD.78 Additionally, the semicircle radius in the electrical impedance spectra (EIS, Fig. 3c) gradually becomes smaller from BTT-BPh to BTT-DBT to BTT-PhPD, indicating the gradually lower charge transfer resistance. Besides, the photocurrent response test shows the gradually higher photocurrent density from BTT-BPh to BTT-DBT to BTT-PhPD (Fig. 3d), which also demonstrates more effective charge separation and migration. These data provide a physical explanation for why the COFs with heteroatom-locked acceptors exhibit higher photocatalytic H2O2 production efficiency: the heteroatom-locked acceptors extend the π-electron delocalization in COFs, which promotes the exciton separation, enhances the transport efficiency of photogenerated charge carriers, and consequently utilizes the absorbed solar energy better.
The above H2O2 reaction intermediate analyses prove the indirect 2 e− ORR process of the three COF-based photocatalytic H2O2 production reactions (Fig. 2c–f), which means the interactions between the COFs and O2-related intermediate play an important role. DFT calculations were conducted to evaluate the oxygen adsorption energies at various sites of the three COFs (Fig. 3e and S18). It can be seen that the oxygen adsorption site of BTT-BPh is situated on the N atom of the imine bond; as for BTT-DBT and BTT-PhPD, the oxygen adsorption sites are located on the S atom of dibenzothiophene and the N atom of phenylpyridine, respectively, suggesting that the incorporation of heteroatoms has altered the O2 adsorption sites. Besides, the oxygen adsorption energy gradually enhances from BTT-BPh (−0.13 eV) to BTT-DBT (−0.16 eV) to BTT-PhPD (−0.21 eV), which confirms that the incorporation of the S/N heteroatom can effectively improve the oxygen adsorption capacity on the COF surface. Additionally, the changes in Gibbs free energies for the reaction intermediates on the COF surfaces were also simulated. As shown in Fig. 3f, the processes O2 → *O2 and *OOH → *HOOH exhibited positive free energy changes (the rate-determining steps), in which the Gibbs free energy gradually become lower from BTT-BPh (3.12 eV and 2.54 eV) to BTT-DBT (2.52 eV and 1.91 eV) to BTT-PhPD (0.84 eV and 1.54 eV); moreover, the energy barriers in the generation of other intermediates (e.g. *OOH and H2O2) also decreased from BTT-BPh to BTT-DBT to BTT-PhPD. These all indicate that the introduction of the N/S-heteroatom has enhanced the surface reaction kinetics of the ORR process toward H2O2 production.
As mentioned above, O2 mass transport is the first and a key step for the surface ORR, due to its low solubility in water (8 mg L−1 at 25 °C and 1 atm) and low diffusion coefficient (2.1 × 10−5 cm−2 s−1). Actually, O2 mass transport not only can be regulated through the structural modulation of the catalyst, but also can be controlled by the reaction system.79 Consequently, to further enhance the O2 diffusion efficiency, a gas diffusion system was introduced into the flow reaction system (Fig. 4a) instead of the above classic testing system (catalyst dispersed in water with stirring in an air atmosphere, Fig. S19) with BTT-PhPD (COF with the best photocatalytic H2O2 production performance) as the catalyst. Different amounts of the catalyst mixed with a binder were attached onto porous carbon paper (details in the SI), which was then placed between transparent quartz plates. Water was continuously flowing over one side of the catalyst-loaded carbon paper, driven by a peristaltic pump, while atmospheric oxygen entered through the other side and diffused into the COF surface to participate in the reaction. Upon illumination, atmospheric O2 and flowing H2O were catalysed by the COF to produce H2O2 at the three-phase interface, and the H2O2 solution was recycled and collected in a liquid storage bottle (Fig. 4a). The H2O2 production rates per reaction area and per unit mass were measured over 4 hours for catalyst loadings of 1 mg, 2 mg, 3 mg, and 5 mg. It was observed that as the loading amount decreased, the H2O2 production rate per unit reaction area slightly decreased (Fig. 4b), whereas the production rate per unit mass significantly increased (Fig. 4c). This may be due to the increased catalyst thickness at higher loadings, which impeded oxygen mass transport and reduced light utilization efficiency. The highest H2O2 production rate (4.06 mmol g−1 h−1) was achieved with 1 mg catalyst loading, showing a significant improvement versus the traditional system (Fig. S20). Furthermore, a decrease in the water flow rate (20 ml min−1 to 2 ml min−1) resulted in a slight enhancement in the photocatalytic H2O2 production efficiency (4.0 mmol g−1 h−1 to 4.7 mmol g−1 h−1) (Fig. S21). Additionally, the catalyst recycling process was very convenient, because the H2O2 solution in the storage bottle and the testing system could be easily replaced with fresh water for the next cycling experiment. As shown in Fig. 4d, the catalyst retained stable H2O2 production rate after five cycles, demonstrating the great stability of the reaction system. Consequently, compared to the traditional reaction system, the introduction of the gas diffusion system not only improves the O2 mass transport, but also realizes the immobilization and efficient recycling of the COF catalyst.
Footnote |
| † Q. Nan and J. Ning contributed equally to this work. |
| This journal is © The Royal Society of Chemistry 2025 |