Open access peer-reviewed chapter

Photocatalysis of Covalent Organic Frameworks

Written By

Hui Liu and Yingjie Zhao

Submitted: 13 August 2022 Reviewed: 02 September 2022 Published: 09 December 2022

DOI: 10.5772/intechopen.107672

From the Edited Volume

Covalent Organic Frameworks

Edited by Yanan Gao and Fei Lu

Chapter metrics overview

256 Chapter Downloads

View Full Metrics

Abstract

The development of clean and sustainable energy is gaining attention in light of the current energy crisis and global warming. An ideal way to utilize renewable solar energy is to convert clean energy through photocatalysis. This includes splitting water, reducing CO2, regenerating coenzymes, etc. Photocatalysis relies heavily on photocatalysts. It has recently become popular to use organic porous polymers in this process. Covalent organic frameworks (COFs), as one of the organic porous polymers, have the characteristics of high crystallinity, porosity, and structural designability that make them perfect platforms for photocatalysis. An overview of recent advances in COF photocatalysts is presented in this chapter. The photocatalytic applications of COFs with different ligation and different structures were first discussed, including photocatalytic hydrogen evolution, CO2 conversion, coenzyme regeneration, and conventional organic reactions. Finally, conclusions and prospects were provided in the last section.

Keywords

  • covalent organic frameworks
  • photocatalysis
  • photocatalytic hydrogen evolution
  • photocatalytic CO2 reduction

1. Introduction

As fossil fuels (e.g. oil, coal, and natural gas) were excessively utilized, energy crisis and climate issues gradually became global concerns [1, 2, 3]. Developing “green energy” as a replacement for traditional fossil fuels will be the most efficient solution to these problems. Photocatalysis is a green technology with important application prospects in energy and environment. Photocatalytic reaction refers to the process of semiconductor photocatalysts promoting the conversion of compounds under light conditions, which can effectively convert light energy into chemical energy. On the one hand, photocatalytic degradation of organic pollutants and photocatalytic reduction of CO2 is helpful to solve environmental problems. On the other hand, photolysis of hydrogen and oxygen in aquatic products can develop new energy for human beings. It should be noted that photocatalyst materials play an important role in the whole photocatalytic process. During the photocatalytic process, electron–hole pairs in the photocatalysts are generated by absorption of photons with higher energy than the band gap. In redox reactions, electrons and holes migrate to the surface, catalyzing the reaction. The first photoelectrochemical water splitting on TiO2 was reported by Fujishima and Honda in 1972 [4]. Since then, a series of inorganic photocatalysts including TiO2 [5, 6], cadmium sulphide (CdS) [7, 8], zinc oxide (ZnO) [9] and silver phosphate (Ag3PO4) [10] emerged and lead the research over the past several decades. Its practical application was limited, however, by the requirements for UV light, heavy metal toxicity, and photo corrosion. In recent years, using organic polymers like linear conjugated polymers (CPs) [11, 12, 13, 14], graphitic carbon nitride (g-C3N4) [15, 16, 17], conjugated microporous polymers (CMPs) [18, 19, 20, 21, 22, 23], covalent organic frameworks (COFs) [24, 25, 26, 27] and covalent triazine-based frameworks (CTFs) [28, 29, 30, 31] as photocatalysts became hot topics. Among them, COFs have drawn wide attention due to advantages of easy structure design, large surface area, tunable electronic properties and band gaps, and diverse synthesis methods.

COFs are a new type of crystalline organic porous polymer based on covalent bond connection [32]. COFs with two-dimensional (2D) or three-dimensional (3D) networks could be prepared according to the construction units of different building blocks. The unique structures of COFs also bring a couple of important advantages: (1) the structural designability endows them with enhanced visible-light absorption and tunable band structure; (2) the porous structure is conducive to the adsorption of substrates and the transport of products; (3) the strong covalent bonds endow COFs with good stability, which thereby prolongs the lifetime of the photoactive structure; (4) the ordered conjugated structure is beneficial not only to the absorption of light energy but also to the transport of excited electrons. As a result of these potential advantages, COFs have been extensively used in gas storage and separation [33, 34, 35, 36], catalysis [37, 38, 39], optoelectronics [40], sensing [41, 42, 43], and energy storage [44, 45]. There have been many excellent reviews about the synthesis, structures, and applications of COFs materials [46, 47, 48, 49, 50, 51, 52, 53, 54, 55]. It can be said that COFs material has many advantages as a heterogeneous photocatalyst.

In this chapter, we will focus on the photocatalytic applications of COFs including water splitting, CO2 reduction, coenzyme regeneration, and photocatalytic organic reactions. Finally, we will discuss the challenges and opportunities of COFs as photocatalysts.

Advertisement

2. Photocatalytic applications of COFs

2.1 Photocatalytic hydrogen evolution

Hydrogen energy is regarded as the most promising clean energy in the 21st century because the only product of hydrogen combustion is water [56]. Hydrogen production from powder photocatalyst is expected to break the cost barrier and become the cheapest technology to decompose water to produce hydrogen, which is expected to surpass fossil fuels. In recent years, organic semiconductor materials have been extensively explored in photocatalytic hydrogen evolution. Among them, COFs with well-ordered conjugate structures showed great advantages not only in photocatalytic performance but also in the deep understanding of the structure–activity relationship [55, 56, 57, 58]. A great deal of research work has focused on improving the performance of the photocatalyst by adjusting the building units and linkages of the COFs. In this section, representative building blocks and linkages of COFs for photocatalytic hydrogen evolution were summarized and discussed.

2.1.1 Hydrazone-linked COFs

Hydrazone-linked COFs with hydrolytic and oxidative stability were prepared by condensation between hydrazide and aldehyde derivatives. Therefore, hydrazone-linked COFs provide a valuable design platform for water-splitting photocatalysts. In 2014, Lotsch group constructed a 2D COF (TFPT-COF) with hydrazone-linked through the condensation reaction of 2,5-diethoxyterephthalohydrazide and triazine-based aldehyde (Figure 1) [25]. This is the first report of COFs materials applied to photocatalytic hydrogen production. TFPT-COF is a mesoporous material with a pore size of 3.8 nm, a specific surface area of 1603 m2/g, and a pore volume of 1.03 cm3/g. Compared with monomer, TFPT-COF has better visible light absorption performance, and the absorption boundary can reach more than 600 nm. TFPT-COF was selected as a photosensitizer, while Pt was used as the proton reduction catalyst. When illuminated with visible light, the hydrogen evolution rate in the first five hours reached 1970 μmol h−1 g−1 using 10 vol% aqueous triethanolamine (TEOA) solution as the sacrificial electron donor. The quantum efficiency was determined to be 2.2%. Under the same conditions, the hydrogen production efficiency of TFPT-COF was much superior to Pt-modified amorphous melon, g-C3N4, and crystalline poly (triazine imide) [59].

Figure 1.

(a) Synthetic route and top view of the TFPT-COF. (b) Photocatalytic performance of Pt-modified TFPT-COF with different SED (black: sodium ascorbate solution; red: triethanolamine solution).

As a type of conjugated macrocycle, porphyrin displays unique photophysical and redox characteristics [60]. Wang and co-workers designed and synthesized four isostructural porphyrin-based 2D COFs (Figure 2) [61]. By incorporating different transition metals into the porphyrin rings, the physical and electronic properties of COFs were rationally tuned. There was a high level of crystallinity and surface area in all of the COFs. When illuminated with visible light while containing Pt as a co-catalyst and TEOA as a sacrificial electron donor, these four COFs exhibited tunable hydrogen evolution efficiency with the order of CoPor-DETH-COF (25 μmol g−1 h−1) < H2Por-DETH-COF (80 μmol g−1 h−1) < NiPor-DETH-COF (211 μmol g−1 h−1) < ZnPor-DETH-COF (413 μmol g−1 h−1). A molecular engineering approach was mainly responsible for the tunable photocatalytic performance. Moreover, hydrazone-linked COFs were also can be converted into oxadiazole-linked COFs via oxidization reaction, which endow COFs with narrow bandgaps and continuous π-electron delocalization for high photocatalytic activity (2615 μmol h−1 g−1) [62].

Figure 2.

Schematic representation of the synthesis of MPor-DETH-COFs.

2.1.2 Azine-linked COFs

The azine-linked COFs are synthesized by the condensation of aldehyde derivatives with hydrazine. The electronic band structure and steric hindrance can be regulated by introducing different heteroatoms into COFs. Lostch and co-workers synthesized a series of azine-linked COFs and investigated the connection between the number of nitrogen atoms in their structures and the photocatalytic hydrogen production efficiency (Figure 3) [63]. Pt and 10 vol% TEOA were selected as the co-catalyst and sacrificial electron donor, respectively. The photocatalytic performance was gradually enhanced with the increase of nitrogen content in the central benzene ring. The hydrogen production efficiencies of N0-COF, N1-COF, N2-COF, and N3-COF were 23, 90, 438, and 1703 μmol h−1 g−1, respectively. Through theoretical calculation, it is found that the planarity of COFs increased with the increase of nitrogen element content. Therefore, N3-COF has the best planarity, which was conducive to the photogenerated electron migration and thus improves the corresponding photocatalytic activity. In addition to the above work, the authors also reported a series of planar pyrene-azine-COFs (A-TEXPY-COFs) with varied nitrogen atoms in the peripheral aromatic units [64]. The photocatalytic efficiency of A-TEXPY-COFs was regulated by nitrogen element content in the peripheral aromatic units. The photocatalytic efficiency exhibited a decreasing tendency with an increasing nitrogen element content. Quantum-chemical calculations suggested that the thermodynamic driving force increased with the decrease of nitrogen content. In addition, Lostch and co-workers also developed cobaloxime as a co-catalyst instead of platinum, a precious metal [65]. The HER rate of N2-COF could reach 782 μmol h−1 g−1 with TEOA as SED.

Figure 3.

(a) The formation of Nx-COF. (b) Photocatalytic H2 evolution of Nx-COFs as photocatalyst over 8 h under visible light irradiation.

2.1.3 β-ketoenamine-linked COFs

The β-ketoenamine-linked COFs could be synthesized through Schiff base reaction and this kind of COFs shows great potential in photocatalytic hydrogen production due to its excellent solvent stabilities of keto−enamine. In 2018, Cooper et al. synthesized two COFs (S-COF and FS-COF) with ultra-high photocatalytic hydrogen evolution activity using benzo-bis(benzothiophene sulfone) moieties (Figure 4) [27]. Pt acted as the co-catalyst, while the sacrificial electron donor was changed to 0.1 M ascorbic acid. The average photocatalytic HER rate of S-COF and FS-COF were measured to be 4.44 mmol h−1 g−1 and 10.1 mmol h−1 g−1, respectively. Both S-COF and FS-COF have sufficient stability under visible light radiation, and FS-COF could work stably for 50 hours under ascorbic acid conditions. It was noted that HER rate of 1.32 mmol h−1 g−1 for FS-COF was obtained even in the absence of Pt. Moreover, higher HER of 16.1 and 16.3 mmol h−1 g−1 were achieved when FS-COF were filled with Eosin Y and WS5F, respectively.

Figure 4.

Synthetic routes for FS-COF and S-COF.

With the introduction of functional organic units in COFs, tunable platforms have become possible. Acetylene as functional group has received significant attention in the field of photocatalysis. For instance, the β-ketoenamine-linked COFs (TP-EDDA and TP-BDDA) were modified with acetylene and diacetylene groups by Thomas et al. (Figure 5) [66]. TP-BDDA has a narrower band gap than TP-EDDA due to its enhanced conjugated structure. Therefore, TP-BDDA showed better activity of photocatalytic hydrogen evolution. With Pt as co-catalyst and 10% TEOA as sacrificial electron donor, TP-BDDA exhibited an HER of 324 μmol h−1 g−1, substantially higher than that of TP-EDDA (30 μmol h−1 g−1). In addition, the optical and electronic properties of β-ketoenamine-linked COFs could be adjusted by slightly changing the structure of their building blocks. Seki and co-workers synthesized a series of β-ketoenamine-linked isoreticular COFs based on 1,3,5-triformylphloroglucinol and 4,4″-diamino-substituted p-terphenyl derivatives [67]. Among these COFs, electron-deficient benzothiadiazole substituted COF (BtCOF150) exhibited better hydrogen evolution efficiency of 750 μmol h−1 g−1.

Figure 5.

Scheme of the synthesis of TP-EDDA and TP-BDDA COFs.

2.1.4 Imine-linked COFs

In COF synthesis, amine linkages are commonly formed by condensation between aldehydes and amines, which are abundant and easily accessible. Various attempts have been made to construct imine-linked COFs for high photocatalytic activity since the 2010 report on photocatalytic H2 production from conjugated poly(azomethine) networks [68].

The introduction of D-A strategy into COFs was beneficial to enhance charge separation and transport ability and thus enhaning photocatalytic activity. In 2020, Wen and co-workers constructed PyTz-COF with a band gap of 2.20 eV using the electron-rich pyrene (Py) and electron-deficient thiazolo[5,4-d]thiazole (Tz) [69]. Overlapping orbitals between Py and Tz units enabled electron transfer and charge separation. With ascorbic acid as SED and Pt as co-catalyst, the hydrogen production rate of PyTz-COF was up to 2072.4 μmol h−1 g−1. Besides, Dong et al. prepared a benzothiadiazole (BT)-based COF (BT-TAPT-COF) using the same strategy [70]. For at least 64 hours, BT-TAPT-COF demonstrated efficient and steady photocatalytic H2 evolution. The maximum HER rate was 949 μmol h−1 g−1 with SED (ascorbic acid) and co-catalyst (Pt). In addition, Chen and co-workers synthesized BT-based COFs that are efficient photocatalysts by introducing halogen moieties into D-A arrangements (Figure 6) [71]. It was found that Py-ClTP-BT-COF with co-catalyst (Pt) and SED (ascorbic acid) has superior HER rate of 8875 μmol h−1 g−1. At 420 nm, Py-ClTP-BT-COF achieved an apparent quantum efficiency (AQE) of 8.45%, which was higher than most reported COF-based photocatalysts at that time. Besides, the HER rate of Py-ClTP-BT-COF could also reach 2200 μmol h−1 g−1 without co-catalyst.

Figure 6.

(a) The formation of Py-XTP-BT-COFs and stacking mode of Py-ClTP-BT-COF. (b) Photocatalytic hydrogen evolution of Py-XTP-BT-COFs under visible light irradiation.

Although imine-linked COFs show the potential of photocatalytic hydrogen production, but their stability and conjugation degree are still their shortcomings, which also restricts the further improvement of their photocatalytic efficiency.

2.1.5 Vinylene-linked COFs

It has been recently discovered that COFs with vinylene-linked have been fabricated via knoevenagel condensation or aldol condensation. As compared to imine, hydrazone, and azine-linked COFs, vinylene-linked COFs show greater conjugation degree. For vinylene-linked COFs, extended conjugation through C=C linkages not only enhances photothermal stability but also enhances absorbance and exciton migration. It was anticipated that vinylene-linked COFs, which combine chemical stability with crystallinity, porosity, and the ability to conjugate, would be highly photoactive. Although the vinylene-linked has many advantages, the poor reversibility of this linkage makes the construction of vinylene-linked very challenging because the bad self-adjusting process is unfavorable for the formation of highly ordered structures.

Jiang and co-workers first synthesized sp2 carbon frameworks by knoevenagel condensation. Then, a strong electron-withdrawing group (3-ethylrho-danine, ERDN) was introduced at the edge of sp2c-COF to obtain D-A heterojunction (Figure 7) [72]. In consequence, visible light was clearly absorbed more widely. As a result of the push-and-pull effect, exciton migration and charge transport were also facilitated. When illuminated with visible light in the presence of Pt and TEOA, sp2 c-COFERDN exhibited an HER of 2.12 mmol g−1 h−1, higher than sp2 c-COF (1.36 mmol g−1 h−1). According to Zhang and co-workers, vinylene-linked COFs have been investigated as effective photocatalytic water splitter [73, 74]. As an example, they designed and synthesized two triazine-cored COFs (g-C18N3-COF and g-C33N3-COF) with unsubstituted olefin linkages [74]. The higher HER rate was observed in g-C18N3-COF (14.6 μmol h−1) and g-C33N3-COF (3.7 μmol) under the same conditions than in imine-linked COF with a similar topology and unit cell parameters. Furthermore, the ordered conjugated structure of g-C18N3-COF yielded a high photocurrent of 45 μA cm−2 at 0.2 V vs. RHE.

Figure 7.

Schematic representation of the synthesis and photocatalytic hydrogen evolution of sp2c-COFERDN.

In order to achieve high photocatalytic performance, effective charge separation and transport must be achieved. Structures built between donors and acceptors are proving to be highly effective. According to Zhao et al., sp2-carbon-linked COFs with periodic D-A structures were shown in a recent study (Figure 8) [75]. Cyano-vinylene linkages are incorporated into an electron-deficient benzobisthiazole structure to form a conjugated structure. It was found that the two-dimensional networks exhibited a strong D-A effect when electron-rich benzotrithiophene was introduced. Taking advantage of the highly conjugated and ordered structure of D-A, BTH-3 displayed an attractive photocatalytic HER of 15.1 mmol h−1 g−1 in the presence of co-catalyst (Pt) and sacrificial electron donor (0.1 M ascorbic acid).

Figure 8.

Synthetic routes and chemical structures of COFs BTH-1, 2, 3.

2.2 Photocatalytic CO2 reduction

Energy crisis and environmental concerns like greenhouse gas emissions can be addressed by converting CO2 into chemical fuels using solar energy [76]. In comparison to photocatalytic hydrogen evolution, CO2 reduction involves more complicated mechanisms. A variety of products could be generated, including carbon monoxide, formic acid, hydrocarbons, and alcohols. As part of the photocatalytic CO2 reduction system, COFs often served as photosensitive supports [77, 78, 79, 80, 81, 82].

A post-synthetic strategy was used by Huang and his colleagues to prepare a re-doped COF for CO2 reduction (Figure 9a) [83]. A triazine-based COF was firstly prepared via reaction between 2,2-bipyridyl-5,5-dialdehyde (BPDA) and 4,4′,4″-(1,3,5-triazine-2,4,6-triyl) trianiline (TTA). After bipyridine ligand was reacted with Re(CO)5Cl, the Re moiety was integrated into COF as a homogeneous photocatalyst for CO2 reduction. Re-COF was found to be an effective photocatalyst for reducing CO2 to CO with high selectivity (98%) (Figure 9b). Because the COFs are crystalline and porous, they demonstrated better activity than homogeneous Re catalysts. Cooper and co-workers combined the rhenium complex with vinyl-linked conjugated Bpy-sp2c-COF affording a Re-Bpy-sp2c-COF heterogeneous photocatalyst [84]. In this case, bipyridine unit offered the ligation to Re complex, was integrated into the vinyl-linked Bpy-sp2c-COF via Knoevenagel condensation. To produce the desired photocatalyst, [Re(CO)5Cl] was loaded on the Bpy-sp2c-COF. In the presence of TEOA, the yield of CO was 1040 μmol g−1 h−1 and the selectivity for H2 was 81% under visible light irradiation. The dye-sensitization process improved selectivity to 86% and CO production rate to1400 μmol g−1 h−1. Furthermore, platinum will generate syngas when added. Moreover, platinum could be tuned to adjust the chemical composition of the gas obtained.

Figure 9.

(a) Synthesis of COF and Re-COF. (b) Proposed Catalytic Mechanism for CO2 reduction.

A synergistic catalyst (Ni-TpBpy) was developed by Zou and co-workers by combining COF (TpBpy) with a single Ni site. [85]. They first constructed a 2,2′-bipyridine-based COF using 5′-diamino-2,2′-bipyridine. Then, the single Ni sites were loaded via treatment with Ni(ClO4)2. In the presence of [Ru(BPY)3]Cl2, the CO yield reached 4057 μmol g−1 within 5 hours, and the selectivity was as high as 96%. A unique microenvironment around single Ni sites in TpBpy was found to be responsible for the excellent photoactivity.

Metalloporphyrin (TAPP) complexes have shown good absorption of visible light and potential to reduce CO2 emissions [86]. Taking that into account, Lan and co-workers constructed a variety of porphyrin-tetrathiafulvalene covalent organic frameworks for CO2 reduction (Figure 10) [87]. By introducing electron-rich tetrathiafulvalene (TTF) molecules into COFs, a donor-acceptor effect was created. Consequently, the generated charge carriers could be efficiently separated and transferred from TTF to TAPP moiety. With TTCOF-Zn as photocatalyst, the yield of CO is 12.33 μmol and the selectivity is close to 100% after 60 hours of visible light irradiation without adding any sacrificant and noble metal cocatalyst. In order to further enhance reduction efficiency, the same group prepared a series of Z-scheme photocatalysts based on COF semiconductors, which combined inorganic semiconductors (TiO2, Bi2WO6, and α-Fe2O3) with COFs (COF-316/318) [88]. It was shown that photocatalytic CO production could reach 69.67 μmol g−1 h−1 without additional photosensitizers and sacrificial agents.

Figure 10.

(a) Schematic of the synthesis of TTCOF-M. (b)Schematic of the mechanism of TTCOF-M CO2RR. (c)Theoretical simulation UV/Vis DRS of TTCOF-Zn and scheme of PET route under light excitation (inset).

Recent studies have shown that COFs without metal active centers can also be used for photocatalytic reduction of carbon dioxide. It is obvious that the reduction efficiency is much less than that of metallic-supplemented COFs, but metal-free COFs are also of great significance for the development of future carbon dioxide reduction. With these concerns, Liu and co-workers prepared a β-ketoenamine-based 2D COF, termed TpBb-COF (Figure 11), by condensation of 2,6-diaminobenzo[1,2-d:4,5-d′]bisthiazole and 1,3,5-triformylphloroglucinol, exhibited an excellent CO production in gas–solid system without using any photosensitizer and sacrificial agent [89]. Interestingly, the reduction of CO2 concentration was more beneficial to CO production, and when CO2 concentration was reduced to 30.0%, the CO generation rate increased to 89.9 μmol g−1 h−1. The mechanism of photocatalytic reduction of carbon dioxide and the rate equation between the production rate of CO and the concentrations of CO2 are given.

Figure 11.

Schematic of the synthesis of TpBb-COF.

In recent years, 3D COFs have also made progress in the field of photocatalytic CO2 reduction. Zhang and co-workers designed eight-linked structural units (TTEP) based on functional porphyrin rings and obtained a 3D COF with pcb topology through Scheff base reaction under solvothermal synthesis conditions (NUST-5 and NUST-6) (Figure 12) [90]. Based on the excellent photoelectric characteristics of porphyrin units, these two COFs have shown great application value in the field of CO2 adsorption and photocatalytic CO2 reduction. After 10 hours of visible light irradiation, the photocatalytic CO2 reduction performance test showed that the CO yield of NUST-5 and NUST-6 was 54.7 μmol g−1 and 76.2 μmol g−1, respectively. In addition, the CH4 yields of NUST-5 and NUST-6 were 17.2 and 12.8 μmol g−1, respectively. The calculated CO/CH4 ratios are 76% and 86%.

Figure 12.

Illustration for the reticular design and synthesis for NUST-5 (R = H) and NUST-6 (R = Me) 3D COFs with pcb topologies.

2.3 Photocatalytic NADH regeneration

As an environmentally friendly and sustainable method of converting solar energy, artificial photosynthesis is modeled after natural photosynthesis. An important component of artificial photosynthesis is coenzyme regeneration (NADH/NADPH) [91]. Many enzyme-mediated synthetic processes require NADH regeneration. Organic polymers with excellent visible light absorption have exhibited excellent photocatalytic performance under the irradiation of visible light [92, 93, 94].

A fully sp2-carbon conjugated COF (TP-COF) was synthesized and characterized for use in photocatalytic NADH regeneration in a previous study by Zhao et al. [95]. TP-COF (Figure 13) was used as a photosensitizer, while [Cp*Rh(bpy)(H)]+ was selected as the electron mediator. Within 10 minutes of irradiation with 420 nm light, the NADH regeneration yield reaches 90.4%. The NADH regeneration system was then coupled with L-glutamate dehydrogenase (GDH) (a redox enzyme) to test the activity of the regenerated NADH. Benefiting from the high efficiency of NADH regeneration, α-ketoglutarate was efficiently converted to L-glutamate with a yield of 97% within 12 min.

Figure 13.

(a) Synthesis and structures of the TP-COF. (b) Illustration of the artificial PSI-induced coenzyme regeneration and photoenzymatic synthesis of l-glutamate by l-glutamate dehydrogenase (GDH).

Also, the linkage effect was investigated in the photocatalytic NADH regeneration [96]. Post-synthesis conversion of two imine-linked COFs (B-COF-1 and T-COF-1) into conjugated COFs (B-COF-2 and T-COF-2) was conducted (Figure 14). As a result of the subtle structure changes, the photoactivity was entirely different. NADH regeneration yielded 74.0% in 10 min with triazine-containing T-COF-2, much higher than with imine-linked precursors. Subsequently, Chen and co-workers introduced bipyridinium units into the framework structure and constructed a mesoporous alkene-linked COF as a porous solid carrier for co-immobilized formate dehydrogenase (FDH) and Rh-group electron medium [97]. By adjusting the incorporation amount of Rh electron medium, it was beneficial to regenerate NADH from NAD+, with an apparent quantum yield (AQY) of 9.17 ± 0.44%. Finally, the assembled photocatalyst-enzyme coupling system can selectively convert CO2 to formic acid, with high efficiency and good reuse.

Figure 14.

Synthesis and Structures of B-COF-1,2; T-COF-1,2.

2.4 Photocatalysis for other organic reactions

The photocatalytic organic reaction has been recognized as a green method for the synthesis of small molecules. In recent years, COFs have been frequently used for photocatalytic synthesis [48, 55, 98]. For instance, Wang and his group reported three COFs (LZU-190, LZU-191, LZU-192) with benzoxazole as the bonding type via a “killing two birds with one stone” strategy (Figure 15) [99]. The COFs structure was kept stable in strong acid and base (9 M NaOH, 9 M HCl), trifluoroacetic acid, boiling water, and light (all for 3 days). Subsequently, it was applied in the experiment of photocatalytic oxidation of arylboronic acids to phenols. Under visible light, all three COFs could oxidize arylboronic acids to phenols using air as oxygen source. Subsequently, LZU-190 was taken as an example to study substrate expansion and mechanism, and electron spin-resonance spectroscopy (ESR) and isotope labeling experiments were used to confirm the mechanism of single electron transfer process.

Figure 15.

One-Pot Construction of Benzoxazole-Linked COFs (LZU-190, LZU-191, and LZU-192) via the Cascade Reactions.

Banerjee and co-workers designed and synthesized a stable COF(TpTt) based on triazine units for photocatalytic E-Z isomerization of olefins (Figure 16) [100]. TpTt has obvious absorption of visible light and can catalyze conversion of trans-stilbene to cis-stilbene under the irradiation of blue light diode. As a result of the good stability of β-ketoenamine linkage, the photoactivity was retained after four cycles. This work also investigated the importance of visible light for this conversion, which begins when light is illuminated and stops when light is closed, suggesting that the reaction takes place through a photocatalytic pathway. In order to further explore the mechanism of photocatalysis, the author uses free radical quenching agent TEMPO for the controllable photocatalytic reaction. When using 4 equiv. TEMPO, cis product yield was significantly reduced to 3%, which confirmed the existence of free radicals and the importance of light catalytic reaction. For different trans olefin substrates, TpTt showed good isomerization of light yield. Besides, Wang et al. developed a red light-driven catalysis system combining the Por-sp2c-COF with TEMPO [101]. When irradiated with 623 nm red light-emitting diode, amines were converted into imines in minutes using this catalytic system.

Figure 16.

Schematic representation of the TpTt synthesis and mechanistic representation of trans to cis photoisomerization of stilbene using the TpTt COF catalyst.

Advertisement

3. Conclusions

In summary, COFs have been viewed as promising materials for photocatalysis applications. As a result of the large specific surface area and the porous structure, light and reactants could be absorbed more efficiently, further optimizing photocatalysis. Conjugated structures with a high degree of order could promote direct charge transport and efficient charge transfer. As an added benefit, COF materials are easily synthesized using the versatile toolbox of organic synthesis. The photoactive building blocks can be altered according to practical applications to synthesize COFs with different structures. A variety of influencing factors, such as pore size, band gap, and other factors, can be adjusted based on the structures. The aforementioned characteristics make COFs a suitable photocatalytic platform. Benefiting from these advantages, COFs have shown excellent capacity in harvesting visible light efficiently and have been frequently used for photocatalytic water splitting, CO2 reduction, coenzyme regeneration, dye degradation, some organic reaction, etc. There are, however, still some problems that need to be solved. It is still challenging to synthesize COFs materials due to the dichotomy of crystallinity and stability. Additionally, the exfoliation of layered 2D COFs during long-term photocatalysis would influence the performance. The exploration of new linkages and synthetic strategies to increase the stability and maintain the crystallinity of the COFs is necessary both for the development of COFs materials and the application of the photocatalyst. Both of these two factors are important to photocatalytic performance. Notably, the application of 3D COFs for photocatalysis is still rare. 3D COFs could avoid the exfoliation problem usually associated with 2D COFs. In addition, the large-scale synthesis of COFs materials is crucial to this field’s future. From the point of view of photocatalysis, the efficiency is still relatively low in such as water splitting and CO2 reduction compared to the traditional inorganic materials. These organic materials should be explored in more detail for some specific applications. Besides, noble metals as co-catalyst and sacrificial agents are normally necessary at the current study lever. It is desirable to avoid the use of noble metals as co-catalyst and sacrificial agents. This could be realized through the structure regulation of the COFs. For example, we could introduce different heteroatom doping (F, Cl, N) and D-A heterojunction structure into structure, which can make the COFs have more reactive active sites and stronger exciton separation and charge transport ability of the COFs material. Furthermore, COFs materials possess highly tunable structures, providing a great advantage for studying photocatalysis mechanisms. By experimenting with different COF catalysts, it is possible to conclude that the structure–activity relationship. We are likely to gain a better understanding of some key points concerning photo-induced charge carriers, such as the separation efficiency and the carrier transport, as time goes on. While traditional photocatalysis has been studied for half a century, the role of COFs materials in photocatalysis is still relatively new. Despite many challenges, we believe COFs materials will offer new development potential in the field of photocatalysis.

Advertisement

Acknowledgments

This work was supported by the National Natural Science Foundation of China (31202117), Natural Science Foundation of Shandong Province (ZR2020ZD38).

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Fouquet R. Path dependence in energy systems and economic development. Nature Energy. 2016;1:16098
  2. 2. Gurney KR, Mendoza DL, Zhou YY, Fischer ML, Miller CC, Geethakumar S, et al. High resolution fossil fuel combustion CO2 emission fluxes for the United States. Environmental Science & Technology. 2009;43:5535-5541
  3. 3. Zhang YY, Yuan S, Feng X, Li HW, Zhou JW, Wang B. Preparation of nanofibrous metal-organic framework filters for efficient air pollution control. Journal of the American Chemical Society. 2016;138:5785-5788
  4. 4. Fujishima A, Honda K. Electrochemical photolysis of water at a semiconductor electrode. Nature. 1972;238:37-38
  5. 5. Li W, Elzatahry A, Aldhayan D, Zhao DY. Core-shell structured titanium dioxide nanomaterials for solar energy utilization. Chemical Society Reviews. 2018;47:8203-8237
  6. 6. Schneider J, Matsuoka M, Takeuchi M, Zhang JL, Horiuchi Y, Anpo M, et al. Understanding TiO2 photocatalysis: Mechanisms and materials. Chemical Reviews. 2014;114:9919-9986
  7. 7. Li Q , Li X, Wageh S, Al-Ghamdi AA, Yu JG. CdS/graphene nanocomposite photocatalysts. Advanced Energy Materials. 2015;5:1500010
  8. 8. Low JX, Dai BZ, Tong T, Jiang CJ, Yu JG. In situ irradiated X-ray photoelectron spectroscopy investigation on a direct Z-scheme TiO2/CdS composite film photocatalyst. Advanced Materials. 2019;31:1802981
  9. 9. Ong CB, Ng LY, Mohammad AW. A review of ZnO nanoparticles as solar photocatalysts: Synthesis, mechanisms and applications. Renewable & Sustainable Energy Reviews. 2018;81:536-551
  10. 10. Martin DJ, Liu GG, Moniz SJA, Bi YP, Beale AM, Ye JH, et al. Efficient visible driven photocatalyst, silver phosphate: Performance, understanding and perspective. Chemical Society Reviews. 2015;44:7808-7828
  11. 11. Sachs M, Sprick RS, Pearce D, Hillman SAJ, Monti A, Guilbert AAY, et al. Understanding structure-activity relationships in linear polymer photocatalysts for hydrogen evolution. Nature. Communications. 2018;9:4968
  12. 12. Shu G, Li YD, Wang Z, Jiang JX, Wang F. Poly(dibenzothiophene-S,S-dioxide) with visible light-induced hydrogen evolution rate up to 44.2 mmol h−1 g−1 promoted by K2HPO4. Applied Catalysis B-Environmental. 2020;261:118230
  13. 13. Sprick RS, Wilbraham L, Bai Y, Guiglion P, Monti A, Clowes R, et al. Nitrogen containing linear poly(phenylene) derivatives for photo-catalytic hydrogen evolution from water. Chemistry of Materials. 2018;30:5733-5742
  14. 14. Tseng PJ, Chang CL, Chan YH, Ting LY, Chen PY, Liao CH, et al. Design and synthesis of Cycloplatinated polymer dots as photocatalysts for visible-light-driven hydrogen evolution. ACS Catalysis. 2018;8:7766-7772
  15. 15. Liu J, Liu Y, Liu NY, Han YZ, Zhang X, Huang H, et al. Metal-free efficient photocatalyst for stable visible water splitting via a two-electron pathway. Science. 2015;347:970-974
  16. 16. Wang G, Zhang T, Yu W, Sun Z, Nie X, Si R, et al. Efficient electronic modulation of g-C3N4 photocatalyst by implanting atomically dispersed Ag1-N3 for extremely high hydrogen evolution rates. CCS Chemistry. 2022;4:2793-2805
  17. 17. Wang XC, Maeda K, Thomas A, Takanabe K, Xin G, Carlsson JM, et al. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nature Materials. 2009;8:76-80
  18. 18. Han CZ, Dong PH, Tang HR, Zheng PY, Zhang C, Wang F, et al. Realizing high hydrogen evolution activity under visible light using narrow band gap organic photocatalysts. Chemical Science. 2021;12:1796-1802
  19. 19. Lee JSM, Cooper AI. Advances in conjugated microporous polymers. Chemical Reviews. 2020;120:2171-2214
  20. 20. Li LW, Cai ZX, Wu QH, Lo WY, Zhang N, Chen LX, et al. Rational design of porous conjugated polymers and roles of residual palladium for photocatalytic hydrogen production. Journal of the American Chemical Society. 2016;138:7681-7686
  21. 21. Sprick RS, Jiang JX, Bonillo B, Ren SJ, Ratvijitvech T, Guiglion P, et al. Tunable organic photocatalysts for visible-light-driven hydrogen evolution. Journal of the American Chemical Society. 2015;137:3265-3270
  22. 22. Yang C, Ma BC, Zhang LZ, Lin S, Ghasimi S, Landfester K, et al. Molecular engineering of conjugated Polybenzothiadiazoles for enhanced hydrogen production by photosynthesis. Angewandte Chemie-International Edition. 2016;55:9202-9206
  23. 23. Zeng QR, Cheng ZH, Yang C, He Y, Meng N, Faul CFJ, et al. Metal-free synthesis of pyridyl conjugated microporous polymers for photocatalytic hydrogen evolution. Chinese Journal of Polymer Science. 2021;39:1004-1012
  24. 24. Ren X, Li C, Kang W, Li H, Ta N, Ye S, et al. Enormous promotion of photocatalytic activity through the use of near-single layer covalent organic frameworks. CCS Chemistry. 2022;4:2429-2439
  25. 25. Stegbauer L, Schwinghammer K, Lotsch BV. A hydrazone-based covalent organic framework for photocatalytic hydrogen production. Chemical Science. 2014;5:2789-2793
  26. 26. Wang J, Zhang J, Peh SB, Liu GL, Kundu T, Dong JQ , et al. Cobalt-containing covalent organic frameworks for visible light-driven hydrogen evolution. Science China-Chemistry. 2020;63:192-197
  27. 27. Wang XY, Chen LJ, Chong SY, Little MA, Wu YZ, Zhu WH, et al. Sulfone-containing covalent organic frameworks for photocatalytic hydrogen evolution from water. Nature Chemistry. 2018;10:1180-1189
  28. 28. Cheng G, Wang KW, Wang SY, Guo LP, Wang ZJ, Jiang JX, et al. Pyrene-based covalent triazine framework towards high-performance sensing and photocatalysis applications. Science China Materials. 2021;64:149-157
  29. 29. Liu MY, Jiang K, Ding X, Wang SL, Zhang CX, Liu J, et al. Controlling monomer feeding rate to achieve highly crystalline covalent triazine frameworks. Advanced Materials. 2019;31:1807865
  30. 30. Wang KW, Yang LM, Wang X, Guo LP, Cheng G, Zhang C, et al. Covalent triazine frameworks via a Low-temperature polycondensation approach. Angewandte Chemie-International Edition. 2017;56:14149-14153
  31. 31. Wang N, Cheng G, Guo LP, Tan BE, Jin SB. Hollow covalent triazine frameworks with variable Shell thickness and morphology. Advanced Functional Materials. 2019;29:1904781
  32. 32. Huang X, Sun C, Feng X. Crystallinity and stability of covalent organic frameworks. Science China-Chemistry. 2020;63:1367-1390
  33. 33. Fan HW, Mundstock A, Feldhoff A, Knebel A, Gu JH, Meng H, et al. Covalent organic framework-covalent organic framework bilayer membranes for highly selective gas separation. Journal of the American Chemical Society. 2018;140:10094-10098
  34. 34. Gao Q , Li X, Ning GH, Xu HS, Liu CB, Tian BB, et al. Covalent organic framework with frustrated bonding network for enhanced carbon dioxide storage. Chemistry of Materials. 2018;30:1762-1768
  35. 35. Kang ZX, Peng YW, Qian YH, Yuan DQ , Addicoat MA, Heine T, et al. Mixed matrix membranes (MMMs) comprising exfoliated 2D covalent organic frameworks (COFs) for efficient CO2 separation. Chemistry of Materials. 2016;28:1277-1285
  36. 36. Li ZP, Zhi YF, Feng X, Ding XS, Zou YC, Liu XM, et al. An azine-linked covalent organic framework: Synthesis, characterization and efficient gas storage. Chemistry-a European Journal. 2015;21:12079-12084
  37. 37. Deng YW, Zou T, Tao XF, Semetey V, Trepout S, Marco S, et al. Poly(epsilon-caprolactone)-block-polysarcosine by ring-opening polymerization of sarcosine N-Thiocarboxyanhydride: Synthesis and Thermoresponsive self-assembly. Biomacromolecules. 2015;16:3265-3274
  38. 38. Liu H, Yan XL, Chen WB, Xie Z, Li S, Chen WH, et al. Donor-acceptor 2D covalent organic frameworks for efficient heterogeneous photocatalytic alpha-oxyamination. Science China-Chemistry. 2021;64:827-833
  39. 39. Liu Y, Wu C, Sun Q , Hu F, Pan Q , Sun J, et al. Spirobifluorene-based three-dimensional covalent organic frameworks with rigid topological channels as efficient heterogeneous catalyst. CCS Chemistry. 2021;3:2418-2427
  40. 40. Keller N, Bein T. Optoelectronic processes in covalent organic frameworks. Chemical Society Reviews. 2021;50:1813-1845
  41. 41. Ding SY, Dong M, Wang YW, Chen YT, Wang HZ, Su CY, et al. Thioether-based fluorescent covalent organic framework for selective detection and facile removal of mercury(II). Journal of the American Chemical Society. 2016;138:3031-3037
  42. 42. Liu XG, Huang DL, Lai C, Zeng GM, Qin L, Wang H, et al. Recent advances in covalent organic frameworks (COFs) as a smart sensing material. Chemical Society Reviews. 2019;48:5266-5302
  43. 43. Wu XW, Han X, Xu QS, Liu YH, Yuan C, Yang S, et al. Chiral BINOL-based covalent organic frameworks for enantioselective sensing. Journal of the American Chemical Society. 2019;141:7081-7089
  44. 44. Lei ZD, Yang QS, Xu Y, Guo SY, Sun WW, Liu H, et al. Boosting lithium storage in covalent organic framework via activation of 14-electron redox chemistry. Nature. Communications. 2018;9:576
  45. 45. Vitaku E, Gannett CN, Carpenter KL, Shen LX, Abruna HD, Dichtel WR. Phenazine-based covalent organic framework cathode materials with high energy and power densities. Journal of he American Chemical Society. 2020;142:16-20
  46. 46. Diercks CS, Yaghi OM. The atom, the molecule, and the covalent organic framework. Science. 2017;355:6328
  47. 47. Ding SY, Wang W. Covalent organic frameworks (COFs): From design to applications. Chemical Society Reviews. 2013;42:548-568
  48. 48. Geng KY, He T, Liu RY, Dalapati S, Tan KT, Li ZP, et al. Covalent organic frameworks: Design, synthesis, and functions. Chemical Reviews. 2020;120:8814-8933
  49. 49. Gui B, Lin GQ , Ding HM, Gao C, Mal A, Wang C. Three-dimensional covalent organic frameworks: From topology design to applications. Accounts of Chemical Research. 2020;53:2225-2234
  50. 50. Hu XL, Li HG, Tan BE. COFs-based porous materials for photocatalytic applications. Chinese Journal of Polymer Science. 2020;38:673-684
  51. 51. Huang N, Wang P, Jiang DL. Covalent organic frameworks: A materials platform for structural and functional designs. Nature Reviews Materials. 2016;1:16068
  52. 52. Kandambeth S, Dey K, Banerjee R. Covalent organic frameworks: Chemistry beyond the structure. Journal of the American Chemical Society. 2019;141:1807-1822
  53. 53. Tao S, Jiang D. Covalent organic frameworks for energy conversions: Current status, challenges, and perspectives. CCS Chemistry. 2021;3:2003-2024
  54. 54. Waller PJ, Gandara F, Yaghi OM. Chemistry of covalent organic frameworks. Accounts of Chemical Research. 2015;48:3053-3063
  55. 55. Wang H, Wang H, Wang ZW, Tang L, Zeng GM, Xu P, et al. Covalent organic framework photocatalysts: Structures and applications. Chemical Society Reviews. 2020;49:4135-4165
  56. 56. Banerjee T, Gottschling K, Savasci G, Ochsenfeld C, Lotsch BV. H-2 evolution with covalent organic framework photocatalysts. Acs Energy Letters. 2018;3:400-409
  57. 57. Wan YY, Wang L, Xu HX, Wu XJ, Yang JL. A simple molecular design strategy for two-dimensional covalent organic framework capable of visible-light-driven water splitting. Journal of the American Chemical Society. 2020;142:4508-4516
  58. 58. Yang Q , Luo ML, Liu KW, Cao HM, Yan HJ. Covalent organic frameworks for photocatalytic applications. Applied Catalysis B-Environmental. 2020;276:119174
  59. 59. Schwinghammer K, Tuffy B, Mesch MB, Wirnhier E, Martineau C, Taulelle F, et al. Triazine-based carbon nitrides for visible-light-driven hydrogen evolution. Angewandte Chemie-International Edition. 2013;52:2435-2439
  60. 60. Ding YB, Zhu WH, Xie YS. Development of ion Chemosensors based on porphyrin analogues. Chemical Reviews. 2017;117:2203-2256
  61. 61. Chen RF, Wang Y, Ma Y, Mal A, Gao XY, Gao L, et al. Rational design of isostructural 2D porphyrin-based covalent organic frameworks for tunable photocatalytic hydrogen evolution. Nature. Communications. 2021;12:1354
  62. 62. Yang S, Lv H, Zhong H, Yuan D, Wang X, Wang R. Transformation of covalent organic frameworks from N-Acylhydrazone to oxadiazole linkages for smooth Electron transfer in photocatalysis. Angewandte Chemie-International Edition. 2022;61:e202115655
  63. 63. Vyas VS, Haase F, Stegbauer L, Savasci G, Podjaski F, Ochsenfeld C, et al. A tunable azine covalent organic framework platform for visible light-induced hydrogen generation. Nature. Communications. 2015;6:8508
  64. 64. Stegbauer L, Zech S, Savasci G, Banerjee T, Podjaski F, Schwinghammer K, et al. Tailor-made photoconductive pyrene-based covalent organic frameworks for visible-light driven hydrogen generation. Advanced Energy Materials. 2018;8:1703278
  65. 65. Banerjee T, Haase F, Savasci G, Gottschling K, Ochsenfeld C, Lotsch BV. Single-site photocatalytic H2 evolution from covalent organic frameworks with molecular Cobaloxime Co-catalysts. Journal of the American Chemical Society. 2017;139:16228-16234
  66. 66. Pachfule P, Acharjya A, Roeser J, Langenhahn T, Schwarze M, Schomacker R, et al. Diacetylene functionalized covalent organic framework (COF) for photocatalytic hydrogen generation. Journal of the American Chemical Society. 2018;140:1423-1427
  67. 67. Ghosh S, Nakada A, Springer MA, Kawaguchi T, Suzuki K, Kaji H, et al. Identification of prime factors to maximize the photocatalytic hydrogen evolution of covalent organic frameworks. Journal of the American Chemical Society. 2020;142:9752-9762
  68. 68. Schwab MG, Hamburger M, Feng X, Shu J, Spiess HW, Wang X, et al. Photocatalytic hydrogen evolution through fully conjugated poly(azomethine) networks. Chemical Communications. 2010;46:8932-8934
  69. 69. Li W, Huang X, Zeng T, Liu YA, Hu W, Yang H, et al. Thiazolo[5,4-d]thiazole-based donor-acceptor covalent organic framework for sunlight-driven hydrogen evolution. Angewandte Chemie-International Edition. 2021;60:1869-1874
  70. 70. Wang G-B, Li S, Yan C-X, Lin Q-Q , Zhu F-C, Geng Y, et al. A benzothiadiazole-based covalent organic framework for highly efficient visible-light driven hydrogen evolution. Chemical Communications. 2020;56:12612-12615
  71. 71. Chen W, Wang L, Mo D, He F, Wen Z, Wu X, et al. Modulating benzothiadiazole-based covalent organic frameworks via halogenation for enhanced photocatalytic water splitting. Angewandte Chemie-International Edition. 2020;59:16902-16909
  72. 72. Jin EQ , Lan ZA, Jiang QH, Geng KY, Li GS, Wang XC, et al. 2D sp(2) carbon-conjugated covalent organic frameworks for photocatalytic hydrogen production from water. Chem. 2019;5:1632-1647
  73. 73. Bi S, Yang C, Zhang WB, Xu JS, Liu LM, Wu DQ , et al. Two-dimensional semiconducting covalent organic frameworks via condensation at arylmethyl carbon atoms. Nature Communications. 2019;10:2467
  74. 74. Wei SC, Zhang F, Zhang WB, Qiang PR, Yu KJ, Fu XB, et al. Semiconducting 2D triazine-cored covalent organic frameworks with unsubstituted olefin linkages. Journal of the American Chemical Society. 2019;141:14272-14279
  75. 75. Wang YC, Hao WB, Liu H, Chen RZ, Pan QY, Li ZB, et al. Facile construction of fully sp2-carbon conjugated two-dimensional covalent organic frameworks containing benzobisthiazole units. Nature. Communications. 2022;13:100
  76. 76. Dhakshinamoorthy A, Navalon S, Corma A, Garcia H. Photocatalytic CO2 reduction by TiO2 and related titanium containing solids. Energy & Environmental Science. 2012;5:9217-9233
  77. 77. Gong YN, Zhong WH, Li Y, Qiu YZ, Zheng LR, Jiang J, et al. Regulating photocatalysis by spin-state manipulation of cobalt in covalent organic frameworks. Journal of the American Chemical Society. 2020;142:16723-16731
  78. 78. Kim YH, Kim N, Seo JM, Jeon JP, Noh HJ, Kweon DH, et al. Benzothiazole-based covalent organic frameworks with different symmetrical combinations for photocatalytic CO2 conversion. Chemistry of Materials. 2021;33:8705-8711
  79. 79. Nguyen HL, Alzamly A. Covalent organic frameworks as emerging platforms for CO2 photoreduction. ACS Catalysis. 2021;11:9809-9824
  80. 80. Skorjanc T, Shetty D, Mahmoud ME, Gandara F, Martinez JI, Mohammed AK, et al. Metallated Isoindigo-porphyrin covalent organic framework photocatalyst with a narrow band gap for efficient CO2 conversion. ACS Applied Materials & Interfaces. 2022;14:2015-2022
  81. 81. Wang XY, Fu ZW, Zheng LR, Zhao C, Wang X, Chong SY, et al. Covalent organic framework nanosheets embedding single cobalt sites for photocatalytic reduction of carbon dioxide. Chemistry of Materials. 2020;32:9107-9114
  82. 82. Yang Y, Lu Y, Zhang HY, Wang Y, Tang HL, Sun XJ, et al. Decoration of active sites in covalent-organic framework: An effective strategy of building efficient photocatalysis for CO2 reduction. ACS Sustainable Chemistry & Engineering. 2021;9:13376-13384
  83. 83. Yang SZ, Hu WH, Zhang X, He PL, Pattengale B, Liu CM, et al. 2D covalent organic frameworks as intrinsic photocatalysts for visible light-driven CO2 reduction. Journal of the American Chemical Society. 2018;140:14614-14618
  84. 84. Fu ZW, Wang XY, Gardner A, Wang X, Chong SY, Neri G, et al. A stable covalent organic framework for photocatalytic carbon dioxide reduction. Chemical Science. 2020;11:543-550
  85. 85. Zhong WF, Sa RJ, Li LY, He YJ, Li LY, Bi JH, et al. A covalent organic framework bearing single Ni sites as a synergistic photocatalyst for selective photoreduction of CO2 to CO. Journal of the American Chemical Society. 2019;141:7615-7621
  86. 86. Kilsa K, Kajanus J, Macpherson AN, Martensson J, Albinsson B. Bridge-dependent electron transfer in porphyrin-based donor-bridge-acceptor systems. Journal of the American Chemical Society. 2001;123:3069-3080
  87. 87. Lu M, Liu J, Li Q , Zhang M, Liu M, Wang JL, et al. Rational design of crystalline covalent organic frameworks for efficient CO2 photoreduction with H2O. Angewandte Chemie-International Edition. 2019;58:12392-12397
  88. 88. Zhang M, Lu M, Lang ZL, Liu J, Liu M, Chang JN, et al. Semiconductor/covalent-organic-framework Z-scheme heterojunctions for artificial photosynthesis. Angewandte Chemie-International Edition. 2020;59:6500-6506
  89. 89. Cui J-X, Wang L-J, Feng L, Meng B, Zhou Z-Y, Su Z-M, et al. A metal-free covalent organic framework as a photocatalyst for CO2 reduction at low CO2 concentration in a gas-solid system. Journal of Materials Chemistry A. 2021;9:24895-24902
  90. 90. Shan Z, Wu M, Zhu D, Wu X, Zhang K, Verduzco R, et al. 3D covalent organic frameworks with interpenetrated pcb topology based on 8-connected cubic nodes. Journal of the American Chemical Society. 2022;144:5728-5733
  91. 91. Wang XD, Saba T, Yiu HHP, Howe RF, Anderson JA, Shi JF. Cofactor NAD(P)H regeneration inspired by heterogeneous pathways. Chem. 2017;2:621-654
  92. 92. Liu J, Huang JH, Zhou H, Antonietti M. Uniform graphitic carbon nitride nanorod for efficient photocatalytic hydrogen evolution and sustained Photoenzymatic catalysis. Acs Applied Materials & Interfaces. 2014;6:8434-8440
  93. 93. Pan QY, Liu H, Zhao YJ, Chen SQ , Xue B, Kan XN, et al. Preparation of N-Graphdiyne nanosheets at liquid/liquid Interface for photocatalytic NADH regeneration. ACS Applied Materials & Interfaces. 2019;11:2740-2744
  94. 94. Wang YC, Liu H, Pan QY, Ding NX, Yang CM, Zhang ZH, et al. Construction of Thiazolo 5,4-d thiazole-based two-dimensional network for efficient photocatalytic CO2 reduction. Acs Applied Materials & Interfaces. 2020;12:46483-46489
  95. 95. Zhao YJ, Liu H, Wu CY, Zhang ZH, Pan QY, Hu F, et al. Fully conjugated two-dimensional sp2-carbon covalent organic frameworks as artificial photosystem I with high efficiency. Angewandte Chemie-International Edition. 2019;58:5376-5381
  96. 96. Wang YC, Liu H, Pan QY, Wu CY, Hao WB, Xu J, et al. Construction of fully conjugated covalent organic frameworks via facile linkage conversion for efficient Photoenzymatic catalysis. Journal of the American Chemical Society. 2020;142:5958-5963
  97. 97. Zhao Z, Zheng D, Guo M, Yu J, Zhang S, Zhang Z, et al. Engineering olefin-linked covalent organic frameworks for Photoenzymatic reduction of CO2. Angewandte Chemie International Edition. 2022;61:e202200261
  98. 98. Chen D, Chen WB, Zhang G, Li S, Chen WH, Xing GL, et al. N-rich 2D heptazine covalent organic frameworks as efficient metal-free photocatalysts. ACS Catalysis. 2022;12:616-623
  99. 99. Wei PF, Qi MZ, Wang ZP, Ding SY, Yu W, Liu Q , et al. Benzoxazole-linked Ultrastable covalent organic frameworks for photocatalysis. Journal of the American Chemical Society. 2018;140:4623-4631
  100. 100. Bhadra M, Kandambeth S, Sahoo MK, Addicoat M, Balaraman E, Banerjee R. Triazine functionalized porous covalent organic framework for photo-organocatalytic E-Z isomerization of olefins. Journal of the American Chemical Society. 2019;141:6152-6156
  101. 101. Shi JL, Chen RF, Hao HM, Wang C, Lang XJ. 2D sp2 carbon-conjugated porphyrin covalent organic framework for cooperative photocatalysis with TEMPO. Angewandte Chemie-International Edition. 2020;59:9088-9093

Written By

Hui Liu and Yingjie Zhao

Submitted: 13 August 2022 Reviewed: 02 September 2022 Published: 09 December 2022