Strategic design of covalent organic frameworks (COFs) for photocatalytic hydrogen generation

Kamal Prakash a, Bikash Mishra b, David Díaz Díaz c, C. M. Nagaraja *a and Pradip Pachfule *b
aDepartment of Chemistry, Indian Institute of Technology Ropar, Rupnagar 140001, Punjab, India. E-mail: cmnraja@iitrpr.ac.in
bDepartment of Chemical and Biological Sciences, S. N. Bose National Centre for Basic Sciences, Kolkata – 700106, India. E-mail: ps.pachfule@bose.res.in
cInstituto Universitario de Bio-Orgánica Antonio González y Departamento de Química Orgánica, Universidad de La Laguna, Avenida Astrofísico Francisco Sánchez, La Laguna 38206, Tenerife, Spain

Received 12th April 2023 , Accepted 19th June 2023

First published on 19th June 2023


Abstract

Covalent organic frameworks (COFs) are an emerging class of crystalline materials that are attracting increasing attention due to their high porosity, crystallinity, and tunable properties. Consequently, the strategic design of COF-based photocatalysts for various applications, including energy and environmental remediation, has attracted considerable interest. In particular, the sustainable production of clean fuel – hydrogen (H2) – by water splitting is a promising means to meet the global energy demand and to address the atmospheric CO2 concentration caused by the excessive use of fossil fuels. In this regard, COFs offer potential advantages due to their modular nature, which facilitates their rational design from suitable organic building blocks to achieve optimal properties of visible light harvesting properties and easy charge transport. As a result, extensive research has been devoted to the design of photoresponsive COFs for efficient water splitting to generate hydrogen. Here, we provide a comprehensive review of recent developments in the strategic design of COF-based photocatalysts for solar fuel production via water splitting. The various organic linkers used in the construction of photocatalytic COFs and their structure–property correlations are discussed in detail. The role of bandgap engineering in tuning the hydrogen evolution efficiency of COFs is also discussed. Furthermore, the current challenges and future perspectives of COF-based solid catalysts for green and sustainable clean fuel production are presented. Indeed, this review demonstrates the importance of COF-based photocatalysts for the visible-light-driven hydrogen evolution reaction (HER) and can be beneficial for the future design of photocatalytic systems.


image file: d3ta02189k-p1.tif

Kamal Prakash

Dr Kamal Prakash received his PhD degree in Chemistry from the Indian Institute of Technology (IIT) Roorkee, India in May 2018. Following his PhD, he worked as a Postdoctoral Fellow at the University of Houston, Texas, USA in 2018–2019. Later, he joined the Indian Institute of Technology Ropar as an Institute Postdoctoral Fellow in 2021. His current research interests focus on the design and synthesis of novel covalent organic frameworks (COFs) and their photocatalytic application in hydrogen evolution and CO2 reduction/conversion.

image file: d3ta02189k-p2.tif

Bikash Mishra

Bikash Mishra is currently pursuing his doctor of philosophy at the department of chemical and biological sciences at S. N. Bose National Centre for Basic Sciences, Kolkata, India. He was awarded his Bachelor of sciences in chemistry from Sidho-Kanho-Birsha University, Purulia. He pursued his master of science in organic chemistry from Banaras Hindu University, Varanasi. Currently, his research focusses on water splitting by covalent organic frameworks (COF) for hydrogen production.

image file: d3ta02189k-p3.tif

David Díaz Díaz

Dr David Díaz Díaz received his PhD in Chemistry from the University of La Laguna (ULL) (2002). Then he joined Prof. Finn's group at TSRI, USA. Since 2006, he has held positions in academia and industry (Ramón y Cajal, UAM, Spain, 2006; Dow, Switzerland, 2007; CSIC, Spain, 2009; University of Regensburg, Germany, Alexander von Humboldt Researcher (2010), Heisenberg Professor (2013), and Privatdozent (since 2018)). In 2020, he was appointed as Distinguished Researcher (ULL). His main research interest focuses on the development of new functional materials for biomedical, environmental, and energy applications.

image file: d3ta02189k-p4.tif

C. M. Nagaraja

Dr C. M. Nagaraja is an associate professor at the Indian Institute of Technology Ropar. He received his PhD from the Indian Institute of Science, Bangalore in 2007. Subsequently, he carried out postdoctoral research at Brandeis University, USA and Jawaharlal Nehru Centre for Advanced Scientific Research (JNCASR), Bangalore. Prior to joining the Indian Institute of Technology Ropar in 2012, he worked as an Assistant Professor at IIT Jodhpur. His research mainly focuses on the design of framework materials for utilization of carbon dioxide and inorganic nanostructured materials for generation of solar fuels.

image file: d3ta02189k-p5.tif

Pradip Pachfule

Dr Pradip Pachfule studied chemistry at Solapur University, India and graduated in 2008. He received his PhD at CSIR-National Chemical Laboratory, Pune, India under the supervision of Prof. Rahul Banerjee in 2014. Later, he worked as a JSPS postdoctoral research fellow in the laboratory of Prof. Qiang Xu at AIST, Kansai, Japan. This was followed by working in the group of Prof. Arne Thomas as an Alexander von Humboldt postdoctoral fellow and a postdoctoral research fellow at the Technische Universität Berlin, Germany (2017–2021). He is currently working as an assistant professor at S. N. Bose National Centre for Basic Sciences, Kolkata, India. His research is focused on covalent organic frameworks and their applications in photocatalytic water splitting and CO2 reduction.


1. Introduction

Since the beginning of industrialization, fossil fuels such as natural gases, petroleum, and coal have been the primary source of energy.1 As a result, the consumption of fossil fuels has rapidly increased over the past few decades, resulting in excessive utilization of these non-renewables. Moreover, this process produces CO2 in large amounts, leading to various environmental issues, such as the greenhouse effect, global warming, and climate change.2 To reduce our dependence on traditional energy sources and to keep the environment safe, it is necessary to decarbonize energy sources and develop carbon-free renewable energy sources such as solar, wind, tidal, biomass, etc. However, most of these energy sources are nondependable.3–7 In these circumstances, hydrogen (H2) has been considered an alternative clean fuel with the potential to be an eco-friendly and cost-effective renewable energy source for future generations.8 Notably, hydrogen is an essential feedstock for many industrial applications, such as petroleum refining, synthesis of fertilizers, and pharmaceuticals. Hydrogen can be accumulated and transported in power plants to electrify gas turbines for electricity generation (Fig. 1).9
image file: d3ta02189k-f1.tif
Fig. 1 The applications and sources of hydrogen. (a) The applications of hydrogen in different fields. (Source: web). (b) Various methods applied for hydrogen generation.

Notably, clean and sustainable hydrogen will be crucial to utilize as synthetic fuel for transportation to prevent greenhouse emissions from fossil fuel-based vehicles. Furthermore, large-scale hydrogen generation can be beneficial because it can be compressed, liquified, and stored for future energy requirements. In addition, green hydrogen is also an important feedstock in the fertilizer and pharmaceutical industries for synthesizing ammonia and other valuable products. Thus, hydrogen is a vital feedstock and a promising environmentally friendly alternative to fossil fuels with zero carbon emissions. However, the large-scale generation, storage, and transportation of hydrogen continues to be a significant challenge that needs to be addressed to realize a complete transition to a hydrogen economy.10–12

Hydrogen can be produced from various resources, such as fossil fuels, biomass, water, agricultural waste, and sewer sludge.13–15 The traditional methods of hydrogen generation from fossil fuels, such as coal and natural gases, are the most dominant, costly, and environmentally hazardous, accounting for the highest CO2 emissions to the atmosphere. However, hydrogen generation from biomass and agricultural waste is a controlled process involving heat, steam, and oxygen without combustion, and as a result, these methods contribute to low-level CO2.16,17 Considering the abundance of natural sunlight and water, the cleanest way to produce hydrogen is through the water-splitting process, in which water molecules are split into hydrogen and oxygen using photocatalysts. This process can be accomplished via photocatalysis using solar energy or photoelectrocatalysis by applying electricity in the presence of solar energy. In a pioneering work, Fujishima and Honda demonstrated hydrogen generation by water splitting using a TiO2 semiconductor as a photocatalyst for the first time.18 Motivated by this work, many metal oxide/sulfide-based semiconductor materials have been studied as photocatalysts for hydrogen generation.19–21 Although metal oxide-based semiconductor nanomaterials such as TiO2, SrTiO3, ZnO, ZrO2, and Fe2O3 have shown great potential due to their high chemical and photostability, most of them are active mainly in the UV region of the solar spectrum, limiting their application under visible light. However, it is worth noting that recently, Domen and coworkers immobilized SrTiO3:Al on a silica support and utilized it for photocatalytic water splitting. This system has shown a solar-to-hydrogen (STH) conversion efficiency of 0.76% per 1 m2 panel, representing the largest demonstration of solar hydrogen generation.22 In addition to metal oxides, metal nitrides and metal sulfide-based semiconductor nanomaterials have also been employed as visible light active photocatalysts for H2 generation.23–31

In addition, layered two-dimensional materials such as graphene oxide (GO), reduced graphene oxide (rGO), and graphitic carbon nitride (g-C3N4) are commonly employed in photocatalysis. Their layered structure promotes facile separation and migration of photoinduced charge carriers, resulting in enhanced photocatalytic efficiency.32 In particular, g-C3N4 is widely studied as a metal-free visible-light-active photocatalyst due to its favorable band gap (∼2.7 eV) and high chemical/thermal stability.33 However, further functionalization of these carbon-based materials and their porosity enhancement are still challenging.34 Furthermore, in contemporary research, porous organic polymers (POPs) have emerged as promising candidates as photocatalysts for hydrogen generation owing to their unique properties of visible light absorption, tailored pore size, and functionalities.35–38 Hence, researchers have targeted various types of porous organic materials, such as hypercross-linked polymers (HCPs),39 polymers of intrinsic microporosity (PIMs),40 porous aromatic frameworks (PAFs),41 covalent triazine frameworks (CTFs),42 and covalent organic frameworks (COFs).43,44 For instance, Yoshino and coworkers investigated the application of poly(p-phenylene) as a photocatalyst for the HER in the presence of trimethylamine (TMA) as a sacrificial electron donor (SED).45 Similarly, Jiang's group designed a conjugated organic polymer encapsulated in a dendritic shell for photocatalytic hydrogen evolution from water.46 Furthermore, the application of a fully conjugated three-dimensional poly(azomethine) for HER has been reported.47 Among these materials, COFs are the most prominent class of porous organic materials with permanent porosity, good thermochemical stability and well-defined crystalline structure.48–53 Owing to their unique properties, COFs offer potential applications in the fields of catalysis, sensing, optoelectronics, energy, drug delivery, and so on.54–56 After the first landmark synthesis of two-dimensional COFs by Yaghi and coworkers43 in 2005, many two-(2D) and three-dimensional (3D) ordered COFs with permanent porosities, large surface areas, and chemical stabilities have been synthesized through covalent bonding between organic precursors.57–59 Additionally, COFs can be structurally predesigned and functionalized by varying their organic synthons as per requirements.60 The chemical and thermal stability of COFs is an added advantage in inhibiting photocorrosion during photocatalysis. Thus, COFs are applied for many photocatalytic transformations, including water splitting to hydrogen and oxygen evolution, CO2 reduction, and organic transformations.61,62

In particular, the rational design of COFs with optimal properties suitable for visible light-promoted water splitting for hydrogen generation has gained significant interest from researchers worldwide. In this direction, intensive research efforts are being made to design COFs, which can promote good visible light absorption, rapid diffusion of charge carriers, and an enhanced lifetime of excited species. The ordered porous network structure creates a good interface with reactants, electrolytes, sacrificial donors, and cocatalysts to render efficient photocatalytic activity. In addition, the extended π-conjugation of organic building blocks can be organized in particular stacking patterns to achieve facile transport of charge carriers, rendering a longer lifetime of excited states. Owing to these unique features, COFs have been identified as highly promising materials for efficient and durable photocatalysts for light-driven water splitting to hydrogen evolution.63,64

2. Basic principle of photocatalytic water splitting

The overall water splitting is a thermodynamically energy-demanding process with a free energy requirement of ΔG = +237.2 kJ mol−1 and comprises simultaneous red-ox processes, as shown below:65
 
image file: d3ta02189k-t1.tif(1)
 
image file: d3ta02189k-t2.tif(2)

The process offers an energy barrier of 1.23 eV that can be overcome by utilizing photocatalysts possessing an optimum band gap that lies in the range of 1.5–3.2 eV.66 For an efficient water-splitting reaction, the conduction band (CB) of the photocatalyst should be more negative than the reduction potential of H+/H2 [NHE (normal hydrogen electrode) at pH = 0], and the valence band (VB) should be more positive than the oxidation potential of O2/H2O (1.23 V vs. NHE at pH = 0).

The basic steps involved in the photocatalytic overall water splitting reaction are depicted in Fig. 2a. The photocatalyst absorbs light with energy higher than the bandgap to generate charge carriers, i.e., electrons and holes at the conduction band (CB) and valence band (VB), respectively. Then, the photogenerated electrons and holes migrate to the catalyst's surface and are utilized for the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) in the presence of reduction and oxidation cocatalysts, respectively. It is important to note that the photogenerated electrons and holes can undergo recombination before being utilized for the redox process, resulting in the loss of charge carriers. Overall water splitting is practically difficult to achieve because it suffers from slow kinetics and unfavorable thermodynamics.67–69 In addition, separating the hydrogen and oxygen generated is another major concern of the overall water-splitting process. Hence, the water-splitting reaction is generally performed as two half-reactions, wherein at one half, the reduction of water to hydrogen evolution and in another half, the oxidation of water to O2 generation is being carried out. Consequently, during hydrogen generation, the oxidation half-reaction is generally replaced by using sacrificial electron donors (SEDs) such as triethanolamine,70 triethylamine,71 ascorbic acid,72 ethylenediaminetetraacetic acid (EDTA),73 and lactic acid (Fig. 2b).74


image file: d3ta02189k-f2.tif
Fig. 2 Mechanisms involved in the water splitting process. (a) Schematic representation of photocatalytic overall water splitting. (b) Photocatalytic hydrogen evolution. (c and d) Type-II and Z-scheme heterojunction systems.

Furthermore, it is important to overcome electron–hole recombination for effective hydrogen generation by water reduction, which is achieved by introducing cocatalyst and heterojunction strategies. Generally, Pt is used as a cocatalyst for COF-based catalysts, which effectively utilize photogenerated electrons to reduce water to hydrogen due to its large work functions and electron-transport properties. Furthermore, heterojunction photocatalysts formed by combining COF with other materials have shown enhanced photocatalytic hydrogen generation aided by facile separation and migration of charge carriers to the interface. Among the various heterojunctions studied, type-II heterojunctions and Z-scheme-based systems are known to be effective in achieving improved HER activity.75,76 As shown in Fig. 2c, in type-II heterojunctions, the CB of photosystem-I (PS-I) is more negative than that of PS-II. This arrangement favors the facile transfer of a photogenerated electron from the CB of PS-I to that of PS-II. However, the holes migrate from the VB of PS-II to PS-I, resulting in effective separation and utilization of electrons for reducing water to hydrogen generation. In the case of Z-scheme heterojunctions, the photoinduced electrons in the CB of PS-II recombine with the holes in the VB of PS-I, resulting in the accumulation of photogenerated electrons on PS-I with a negative CB potential and holes on PS-II with a positive VB potential (Fig. 2d). Thus, electron–hole recombination is prevented, leading to enhanced photocatalytic activity. Overall, the cocatalyst and heterojunction strategies effectively promote efficient separation and migration of the photogenerated charge carriers for improved hydrogen generation efficiency.

3. Construction of covalent organic frameworks for hydrogen generation

COFs – exclusively covalently bonded in a thermodynamically stable ordered form – are constructed using a bottom-up approach from organic linkers through reversible reactions (Fig. 3a). The application of COFs as heterogeneous catalysts for hydrogen generation is growing steadily with time, which is evident from the increased number of articles published over the past few years (Fig. 3b). Notably, in the last few years, there has been an upsurge in research on the design of COF-based photocatalysts for hydrogen evolution, which indicates that COFs are preferred as potential photocatalytic materials for HER owing to their unique properties (Table 1).
image file: d3ta02189k-f3.tif
Fig. 3 (a) Pictorial representation of the general synthesis of COFs (C3 + C2). (b) Year wise number of publications of COF for photocatalytic hydrogen evolution (Source: Clarivate).
Table 1 Early reports of COFs for photocatalytic water splitting
Photocatalyst Light source Cocatalyst Sacrificial agent Hydrogen evolution rate (μmol h−1 g−1) AQY (%) Reference
Early reported covalent organic framework for hydrogen evolution
TFPT-COF Visible light Pt Triethanolamine 1970 2.2 (420 nm) Chem. Sci., 2014, 5, 2789
[thin space (1/6-em)]
Early modified covalent organic frameworks for hydrogen evolution
CdS-COF (inorganic) 300 W Xe lamp Pt Lactic acid 3678 4.2 (420 nm) Chem.–Eur. J., 2014, 20, 15961
CN-COF (organic) 300 W Xe lamp Pt Triethanolamine 10[thin space (1/6-em)]100 20.7 (425 nm) Chem. Commun., 2019, 55, 5829
[thin space (1/6-em)]
Highest hydrogen evolution activity for covalent organic framework
CYANO-CON 300 W Xe lamp Pt Ascorbic acid 134[thin space (1/6-em)]200 82.6 (450 nm) Nat. Commun., 2022, 13, 2357
[thin space (1/6-em)]
Higher apparent quantum efficiency for covalent organic framework
CYANO-CON 300 W Xe lamp Pt Ascorbic acid 134[thin space (1/6-em)]200 82.6 (450 nm) Nat. Commun., 2022, 13, 2357
[thin space (1/6-em)]
Photocatalytic hydrogen evolution activity without cocatalyst
CYANO-CON 300 W Xe lamp Ascorbic acid 17.4 2.72 (450 nm) Nat. Commun., 2022, 13, 2357
PhBp-CTF-Ir Visible light Triethanolamine 242 2.36 (420 nm) ACS Appl. Energy Mater., 2022, 5, 7473
BT-TAPT COF Visible light Ascorbic acid 9.45 Chem. Commun., 2020, 56, 12612
Py-ClTP-BT-COF Visible light Ascorbic acid 2200 Angew. Chem. Int. Ed., 2020, 59, 16902
TpPa-COF-(CH3)2 300 W Xe lamp Sodium ascorbate 72.10 ChemCatChem, 2019,10, 1002
Cu-sulphen-HDCOF-NS Visible light Triethanolamine 36[thin space (1/6-em)]990 5.77 (420 nm) J. Mater. Chem. A,2020, 8, 25094
[thin space (1/6-em)]
Photocatalytic hydrogen evolution activity with molecular cocatalyst
N2-COF 300 W Xe lamp Chloro(pyridine)cobaloxime(Co-I) Triethanolamine 782 0.16 (400 nm) J. Am. Chem. Soc., 2017, 139, 16228
15-[Ni(pymt)2]n/CTF-HC2 300 W Xe lamp [Ni(pymt)2] Triethanolamine 3472 ChemCatChem, 2021, 13, 958
TpDTz-COF 300 W Xe lamp NiME Triethanolamine 941 0.20 (400 nm) J. Am. Chem. Soc., 2019, 141, 11082
TpPa-1-COF 300 W Xe lamp Ni12P5 L-Ascorbic acid 31.6 Small, 2022, 2, 201340
[thin space (1/6-em)]
Covalent organic frameworks for simultaneous hydrogen and oxygen evolution
r-CTF NSs 300 W Xe lamp Pt Triethanolamine HER: 10[thin space (1/6-em)]200 11.3 (420 nm) Angew. Chem. Int. Ed., 2021, 60, 25381
300 W Xe lamp Co AgNO3 OER: 247 5.6 (420 nm)
sp2-COF 300 W Xe lamp Pt Triethanolamine HER:1360 Chem, 2019, 5, 1632
300 W Xe lamp Co AgNO3 OER: 22
CTF-HUST-A1-tBuOK Visible light NiPx + Pt HER: 25.4 0.8 (420 nm) Angew. Chem. Int. Ed., 2020, 59, 6007
Visible light NiPx + Pt OER: 12.9
[thin space (1/6-em)]
Early azine-based covalent organic framework for hydrogen evolution
N3-COF 300 W Xe lamp Pt Triethanolamine 1703 0.44 (450 nm) Nat. Commun., 2015, 6, 8508
[thin space (1/6-em)]
Early imine-based covalent organic framework for hydrogen evolution
CN-COF 300 W Xe lamp Pt Triethanolamine 10[thin space (1/6-em)]100 20.7 (425 nm) Chem. Commun., 2019, 55, 5829
[thin space (1/6-em)]
Early olefin-linked covalent organic framework for hydrogen evolution
sp2-COF 300 W Xe lamp Pt Triethanolamine 1360 Chem, 2019, 5, 1632
sp2-COFERDN 2120 0.4 (495 nm)
[thin space (1/6-em)]
Early hydrazone-based covalent organic framework for hydrogen evolution
TFPT-COF Visible light Pt Triethanolamine 1970 2.2 (420 nm) Chem. Sci., 2014, 5, 2789
[thin space (1/6-em)]
Early β-ketoenamine-based covalent organic framework for hydrogen evolution
TP-BDDA-COF 300 W Xe lamp Pt Triethanolamine 324 1.8 (520 nm) J. Am. Chem. Soc., 2018, 140, 1423
[thin space (1/6-em)]
Early imide-based covalent organic framework for hydrogen evolution
COF-imide Visible light Pt Triethanolamine 34 Adv. Sci., 2020, 1, 902988


A snapshot of the important COFs developed over the past few years with their photocatalytic HER performance is shown in Fig. 4. In a pioneering work, Lotsch and coworkers demonstrated the application of TFPT-COF as a visible light photocatalyst for hydrogen generation from water using Pt as a proton reduction catalyst.77 The COF was synthesized from 1,3,5-tris-(4-formyl-phenyl)triazine (TFPT) and 2,5-diethoxy terephthalohydrazide (DETH) linkers (Fig. 4). The triazine-based TFPT COF featuring mesopores of 3.8 nm in diameter and a high surface area of 1603 m2 g−1 showed a hydrogen evolution rate of 1970 μmol h−1 g−1 (see below). Since then, researchers worldwide have been making extensive research efforts on the rational design of COF materials for hydrogen generation applications.


image file: d3ta02189k-f4.tif
Fig. 4 Representative examples of COFs with their hydrogen evolution performances showcasing the advancement in the hydrogen evolution activity with time.

In an important demonstration, Lotsch and coworkers have shown that the electronic and steric variations in the precursor molecules are transferred to the resulting frameworks, thus leading to progressively enhanced hydrogen generation with an increase in the number of nitrogen atoms in the frameworks.78 Furthermore, the same group demonstrated single-site hydrogen evolution by applying azine-linked N2-COF as a photosensitizer and chloro(pyridine)cobaloxime as a cocatalyst. This is a promising step toward designing “COF-molecular cocatalyst”-based photocatalytic systems free of noble metals (e.g., Pt).79 In 2018, Cooper and coworkers demonstrated the design of sulfone-based COFs named S-COF (S = sulfone) and FS-COF (FS = fused sulfone). As a result of high crystallinity, strong visible light absorption, and the presence of wettable and hydrophilic mesopores in the COF matrix, S-COF and FS-COF showed hydrogen evolution with steady rates of 4.44 and 10.1 mmol h−1 g−1 under visible light (see below).80 Similarly, using halogenated (TaPa-Cl2)81 and donor–acceptor (BT-TAPT)82 type COFs, enhanced photocatalytic hydrogen evolution has been achieved. These results highlight the influence of the electronic and photophysical properties of the COFs on the HER performance. To study the application of fully conjugated 2D COFs, Jiang and coworkers developed sp2-carbon conjugated COFs for water splitting (Fig. 4).140 Owing to its excellent chemical (water, acid, and base) stability and extended conjugation, sp2c-COF showed good HER activity with a rate of 1360 μmol h−1 g−1 under visible light (≥420 nm).

Recently, Thomas and coworkers identified the effect of protonation as a promising means to improve the photocatalytic hydrogen evolution performance of a 2D imine-linked COF.83 The protonated TtaTfa COF showed hydrogen generation at a rate of 20.7 mmol g−1 h−1 under visible light in the presence of ascorbic acid as a sacrificial electron donor (see below). In a similar attempt to improve the photocatalytic hydrogen evolution performance of COFs, incorporation of a β-ketene-cyano donor–acceptor pair in a cyano-functionalized COF nanosheet (CYANO-CON) was achieved by Yang and coworkers.84 These nanosheets have shown an apparent quantum yield (AQY) of up to 82.6% (450 nm) and very high hydrogen generation performance. Inspired by the promising performance of COFs, several photocatalytic COFs have been rationally prepared, and their optical and catalytic properties have been engineered to achieve improved hydrogen evolution. However, further advancement in the design of COFs would be desired to accomplish enhanced hydrogen evolution performance with high durability. To improve the HER activity of COFs, various strategies, such as bandgap engineering, cocatalysts, and heterojunction formation, have been employed, which are discussed in the following sections.

3.1 Favorable properties of COFs as photocatalysts

A material that acts as an efficient photocatalyst should possess the essential requirements of an optimal bandgap, improved light harvesting, high porosity, and crystallinity.85 The crystalline nature of photocatalysts with fewer defect sites is highly desired to achieve enhanced catalytic activity by eliminating recombination of photogenerated charge carriers. The modular nature of these materials allows the strategic design of COFs with tunable pore structure and functionality by the judicious choice of organic building blocks. Thus, one can rationally construct COF-based photocatalysts incorporating multiple functionalities to achieve desired properties such as visible light absorption, facile charge transport, high chemical and photostability along with high porosity and crystallinity (Fig. 5). As a result, COFs offer potential advantages as photocatalysts for hydrogen generation from water, which is also reflected in the rapid increase in the application of COF-based photocatalysts for water splitting (Table 2).
image file: d3ta02189k-f5.tif
Fig. 5 Schematic representation showing essential characteristics of an ideal photocatalyst.
Table 2 The photocatalytic hydrogen evolution properties of functionalized COFs
Name of COF Linker Light intensity Conc. of cocatalyst Sacrificial donor AQY (%) Hydrogen evolution rate (μmol h−1 g−1) Reference
Triazine-based COFs
TFTP-COF 2,5-Diethoxy-terephthalohydrazide (DETH) 300 W Xe lamp, λ ≥ 420 nm 2.4 μL of 8 wt% Pt in H2O 1 mL TEOA 2.2 1970 Chem. Sci., 2014, 5, 2789
N0-COF N0-aldehyde 300 W Xe lamp, λ ≥ 420 nm 2.5 μL of 8.0 wt% Pt in H2O 1 μL TEOA 0.001–0.44 23 Nat. Comm., 2015, 6, 8508
N1-COF N1-aldehyde 300 W Xe lamp, λ ≥ 420 nm 2.5 μL of 8.0 wt% Pt in H2O 1 μL TEOA 0.001–0.44 90 Nat. Comm., 2015, 6, 8508
N2-COF N2-aldehyde 300 W Xe lamp, λ ≥ 420 nm 2.5 μL of 8.0 wt% Pt in H2O 1 μL TEOA 0.001–0.44 438 Nat. Comm., 2015, 6, 8508
N3-COF N3-aldehyde 300 W Xe lamp, λ ≥ 420 nm 2.5 μL of 8.0 wt% Pt in H2O 1 μL TEOA 0.001–0.44 1703 Nat. Comm., 2015, 6, 8508
BT-TAPT-COF 4,40-(Benzothiadiazole-4,7-diyl)dibenzaldehyde (BT) 300 W Xe lamp, λ ≥ 420 nm 10 μL of 8.0 wt% Pt in H2O 0.1 M ascorbic acid 0.19 949 Chem. Commun., 2020, 56, 12612
BDF-TAPT-COF 4,4′-(Benzo[1,2-b:4,5-b′]difuran-4,8-diyl)dibenzaldehyde (BDF-CHO) 300 W Xe lamp, AM 1.5 G 10 μL of 8.0 wt% Pt in H2O 0.1 M ascorbic acid 7.8 1390 Chem. Commun., 2021, 57, 4464
TA-COF Terephthalaldehyde 300 W Xe lamp, full wavelength 3.0 wt% Pt 10 vol% TEOA 10 Appl. Surface Sci., 2021, 537, 148082
TFA-COF 2,3,5,6-Tetrafluoroterephthaldehyde 300 W Xe lamp, full wavelength 3.0 wt% Pt 10 vol% TEOA 80 Appl. Surface Sci., 2021, 537, 148082
TtaTfa_AC 1,3,5-Tris(4-formylphenyl)benzene 300 W Xe lamp, λ ≥ 420 nm 8.0 wt% Pt Ascorbic acid 1.43 20[thin space (1/6-em)]700 Angew. Chem. Int. Ed., 2021, 60, 19797
TpaTfa_AC Tris(4-formylphenyl)amine (Tfa) 300 W Xe lamp, λ ≥ 420 nm 8.0 wt% Pt Ascorbic acid 14[thin space (1/6-em)]900 Angew. Chem. Int. Ed., 2021, 60, 19797
TtaTpa_AC 1,3,5-Tris(4-formylphenyl)benzene (Tpa-CHO) and 1,3,5-tris(4-aminophenyl)benzene 300 W Xe lamp, λ ≥ 420 nm 8.0 wt% Pt Ascorbic acid 10[thin space (1/6-em)]800 Angew. Chem. Int. Ed., 2021, 60, 19797
TTI-COF 2,4,6-Tris(4-aminophenyl)-1,3,5-triazine 300 W Xe lamp, λ > 420 nm 3.0 wt% Pt 0.1 M ascorbic acid 460 ACS Catal., 2022, 12, 10718–10726
TTV-COF 2,4,6-Tris[4-(diethyl-phosphono-methyl)phenyl]-1,3,5-triazine 300 W Xe lamp, λ > 420 nm 3.0 wt% Pt 0.1 M ascorbic acid 5500 ACS Catal., 2022, 12, 10718–10726
TTAN-COF 2,4,6-Tris(4-cyano-methyl-phenyl)-1,3,5-triazine 300 W Xe lamp, λ > 420 nm 3.0 wt% Pt 0.1 M ascorbic acid 11[thin space (1/6-em)]940 ACS Catal., 2022, 12, 10718–10726
NKCOF-112 M [2,2′-Bipyridine]-5,5′-dicarbaldehyde 300 W Xe lamp, λ > 420 nm 50 mM H2PtCl6 TEOA 2800 Adv. Sci., 2022, 2, 203832
NKCOF-113 M 5,5′-Bis-(cyanomethyl)-2,2′-bipyridine 300 W Xe lamp, λ > 420 nm 50 mM H2PtCl6 TEOA 56 13[thin space (1/6-em)]100 Adv. Sci., 2022, 2203832
NKCOF-114 M 5,5′-Bis-(cyanomethyl)-2,2′-bipyridine 300 W Xe lamp, λ > 420 nm 50 mM H2PtCl6 TEOA 1160 Adv. Sci., 2022, 2, 203832
BTH-1 Benzobisthiazole 300 W Xe lamp, λ > 420 nm 8.0 wt% Pt 0.1 M ascorbic acid 1.925 10[thin space (1/6-em)]500 Nat. Commun., 2022, 13, 100
BTH-2 Benzobisthiazole 300 W Xe lamp, λ > 420 nm 8.0 wt% Pt 0.1 M ascorbic acid 0.241 1200 Nat. Commun., 2022, 13, 100
BTH-3 Benzobisthiazole 300 W Xe lamp, λ > 420 nm 8.0 wt% Pt 0.1 M ascorbic acid 1.256 15[thin space (1/6-em)]100 Nat. Commun., 2022, 13, 100
COF-JLU100 1,4-Phenylene-diacetonitrile 300 W Xe lamp, λ > 420 nm 12.0 wt% Pt TEOA 5.13 107[thin space (1/6-em)]380 Angew. Chem. Int. Ed., 2022, 61, e202208919
TThB-TZ-COF Trithienyl-benzene 300 W Xe lamp, λ > 420 nm 8.0 wt% Pt 0.1 M ascorbic acid 0.31 1030 ACS Appl. Mater. Interfaces, 2023, 15, 16794–16800
BTTh-TZ-COF Benzotri-thiophenes 300 W Xe lamp, λ > 420 nm 8.0 wt% Pt 0.1 M ascorbic acid 1.38 5220 ACS Appl. Mater. Interfaces, 2023, 15, 16794–16800
COF-JLU35 Benzo[1,2-b:3,4-b′:5,6-b′′]trithiophene-2,5,8-tricarbaldehyde and 1,4-phenyldiacetoni rile 300 W Xe lamp, λ ≥ 420 nm 1.0 wt% Pt Ascorbic acid 3.21 70[thin space (1/6-em)]800 J. Am. Chem. Soc., 2023, 145, 8364–8374
COF-JLU36 1,4-Phenylene-diamine 300 W Xe lamp, 420 < λ < 780 nm 1.0 wt% Pt 0.1 M ascorbic acid 0.37 23[thin space (1/6-em)]600 J. Am. Chem. Soc., 2023, 145, 8364–8374
COF-OH-0 1,3,5-Tri-formylbenzene 300 W Xe lamp, λ ≥ 420 nm 1.0 wt% Pt Ascorbic acid 110 J. Mater. Chem. A, 2022, 10, 24620–24627
COF-OH-1 2,4,6-Triformyl-phenol 300 W Xe lamp, λ ≥ 420 nm 1.0 wt% Pt Ascorbic acid J. Mater. Chem. A, 2022, 10, 24620–24627
COF-OH-2 2,4,6-Triformyl-resorcinol 300 W Xe lamp, λ ≥ 420 nm 1.0 wt% Pt Ascorbic acid 2910 J. Mater. Chem. A, 2022, 10, 24620–24627
COF-OH-3 2,4,6-Triformyl-phloroglucinol 300 W Xe lamp, λ ≥ 420 nm 1.0 wt% Pt Ascorbic acid 0.15 9890 J. Mater. Chem. A, 2022, 10, 24620–24627
[thin space (1/6-em)]
Pyrene-based COFs
A-TEBPY-COF Hydrazine 300 W Xe lamp, AM 1.5 G 6 μL of 8 wt% Pt in H2O 1 mL TEOA 98 Adv. Energy Mater., 2018, 8, 1703278
A-TENPY-COF Hydrazine 300 W Xe lamp, AM 1.5 G 6 μL of 8 wt% Pt in H2O 1 mL TEOA 22 Adv. Energy Mater., 2018, 8, 1703278
A-TEPPY-COF Hydrazine 300 W Xe lamp, AM 1.5 G 6 μL of 8 wt% Pt in H2O 1 mL TEOA 6 Adv. Energy Mater., 2018, 8, 1703278
sp2c-COFEDRN p-Phenylenediacetonitrile (PDAN) 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt 10 vol% TEOA 2120 Chem, 2019, 5, 1632
sp2c-COF p-Phenylenediacetonitrile (PDAN) 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt 10 vol% TEOA 0.46 1360 Chem, 2019, 5, 1632
Tz-COF-4 Naphthalene-2,6-diamine 300 W Xe lamp, λ ≥ 420 nm 8.0 wt% Pt 0.1 M ascorbic acid 1.3 4296 J. Am. Chem. Soc., 2020, 142, 11131
IL-COF-2 Naphthalene-2,6-diamine 300 W Xe lamp, λ ≥ 420 nm 8.0 wt% Pt 0.1 M ascorbic acid 1644 J. Am. Chem. Soc., 2020, 142, 11131
Py-ClTP-BT COF 4,7-Dibromo-5,6-dichlorobenzo[c][1,2,5]thiadiazole 300 W Xe lamp, λ ≥ 420 nm 5.0 wt% Pt 0.1 M ascorbic acid 8.45 8875 Angew. Chem. Int. Ed., 2020, 59, 16902
Py-FTP-BT COF 4,7-Dibromo-5,6-difluorobenzo[c][1,2,5]thiadiazole 300 W Xe lamp, λ ≥ 420 nm 5.0 wt% Pt 0.1 M ascorbic acid 2875 Angew. Chem. Int. Ed., 2020, 59, 16902
Py-HTP-BT COF 4,7-Dibromo-benzo[c][1,2,5]thiadiazole 300 W Xe lamp, λ ≥ 420 nm 5.0 wt% Pt 0.1 M ascorbic acid 1078 Angew. Chem. Int. Ed., 2020, 59, 16902
NK-COF-108 Benzothiadiazole moieties 300 W Xe lamp, λ ≥ 420 nm 52 μL of 50 mM Pt in H2O 0.1 M ascorbic acid 2.96 11[thin space (1/6-em)]600 ACS Catal., 2021, 11, 2098
PyTz-COF 4,4′-(Thiazolo[5,4-d]thiazole-2,5-diyl)dibenzaldehyde (TzDA) 300 W Xe lamp, AM 1.5 G 3.0 wt% Pt 0.1 M ascorbic acid 2072.4 Angew. Chem. Int. Ed., 2021, 60, 1869
PY-DHBD-COF 1,4-Dihydroxybenzidine 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Ascorbic acid 8.4 71[thin space (1/6-em)]160 Nat. Comm., 2022, 13, 1355
Ni-Bn-COF Bis(1,2-diphenyl-glyoximato)-nickel(II)(Ni(dpg)2) 300 W Xe lamp, λ > 420 nm Pt Ascorbic acid 1805 Angew. Chem. Int. Ed., 2022, 61, e202204326
Ni-Py-COF Bis(1,2-diphenyl-glyoximato)nickel(II) 300 W Xe lamp, λ ≥ 420 nm Pt Ascorbic acid 4.28 13[thin space (1/6-em)]231 Angew. Chem. Int. Ed., 2022, 61, e202204326
DABT-Py-COF 4,4′,4′′,4′′′-(Benzo[c][1,2,5]thiadiazole-4,7-diylbis(9,9-dimethyl-9,10-dihydroacridine-10,2,7-triyl))tetrabenzaldehyde 300 W Xe lamp, AM 1.5 G Pt 0.1 M ascorbic acid 5458 J. Mater. Chem. A, 2023, 11, 4007–4012
Co/Zn-Salen-COF Ethylenediamine 300 W Xe lamp, λ ≥ 420 nm CoN2O2 Ascorbic acid 1378 Angew. Chem. Int. Ed., 2023, 62, e202214143
TAPFy-PhI COF Phthalimide 300 W Xe lamp, λ > 420 nm 1.0 wt% Pt 0.1 M ascorbic acid 3.34 2718 ACS Appl. Mater. Interfaces, 2023, 15, 20310–20316
[thin space (1/6-em)]
Porphyrin-based COFs
H2Por-DETH-COF 2,5-Diethoxy-terephthalohydrazide (DETH) 300 W Xe lamp, λ ≥ 400 nm 2.5 μL of 8 wt% Pt in H2O 50 μL TEOA 80 Nat. Comm., 2021, 12, 1354
CoPor-DETH-COF Co + 2,5-diethoxy-terephthalohydrazide (DETH) 300 W Xe lamp, λ ≥ 400 nm 2.5 μL of 8 wt% Pt in H2O 50 μL TEOA 25 Nat. Commun., 2021, 12, 1354
NiPor-DETH-COF Ni + 2,5-diethoxy-terephthalohydrazide (DETH) 300 W Xe lamp, λ ≥ 400 nm 2.5 μL of 8 wt% Pt in H2O 50 μL TEOA 211 Nat. Commun., 2021, 12, 1354
ZnPor-DETH-COF Zn + 2,5-diethoxy-terephthalohydrazide (DETH) 300 W Xe lamp, λ ≥ 400 nm 2.5 μL of 8 wt% Pt in H2O 50 μL TEOA 0.32 413 Nat. Commun., 2021, 12, 1354
[Mo3S13]2−@ZnP-Pz-DHTP-COF Pyrazine-2,5-dialdehyde and 2,5-dihydroxy-terephthalaldehyde 300 W Xe lamp, λ > 420 nm [Mo3S13]2− Lactic acid 4700 Nat. Commun., 2023, 14, 329
[Mo3S13]2−@ZnP-Pz-PEO-COF Pyrazine-2,5-dialdehyde and 2,5-dihydroxy-terephthalaldehyde 300 W Xe lamp, λ > 420 nm [Mo3S13]2− Lactic acid 5.7 11[thin space (1/6-em)]000 Nat. Commun., 2023, 14, 329
[Mo3S13]2−@ZnP-Pz-COF Pyrazine-2,5-dialdehyde 300 W Xe lamp, λ > 420 nm [Mo3S13]2− Lactic acid 10.1 Nat. Commun., 2023, 14, 329
TP-COF 1,1,2,2-Tetra(4-formyl-(1,1′-biphenyl)ethene 300 W Xe lamp, λ > 420 nm 5.0 wt% Pt 20 vol% TEOA 58.4 Applied Surface Science, 2023, 613, 155966
[thin space (1/6-em)]
1,3,5-Triformylphloroglucinol-containing COFs
TP-DTP 4,4′′-Diamino-p-terphenyl 300 W Xe lamp, λ ≥ 395 nm 3.0 wt% Pt 4 mL (10% V) TEOA 20 J. Am. Chem. Soc., 2018, 140, 1423
TP-EDDA 4,4′-(Ethyne-1,2-diyl)dianiline (EDDA) 300 W Xe lamp, λ ≥ 395 nm 3.0 wt% Pt 4 mL (10% V) TEOA 30 J. Am. Chem. Soc., 2018, 140, 1423
TP-BDDA 4,4′-(buta-1,3-diyne-1,4-diyl)dianiline (BDDA) 300 W Xe lamp, λ ≥ 395 nm 3.0 wt% Pt 4 mL (10% V) TEOA 1.8 324 J. Am. Chem. Soc., 2018, 140, 1423
TP-COF 4,4′′-Diamino-p-terphenyl 300 W Xe lamp, λ ≥ 395 nm 8.0 wt% Pt 0.1 M ascorbic acid 1600 Nat. Chem., 2018, 10, 1180
S–COF 3,7-Diaminodibenzo-[b,d]thiophene sulfone (SA) 300 W Xe lamp, λ ≥ 395 nm 8.0 wt% Pt 0.1 M ascorbic acid 4440 Nat. Chem., 2018, 10, 1180
FS-COF 3,9-Diamino-benzo[1,2-b:4,5-b′]-bis[1]benzothiophene sulfone (FSA) 300 W Xe lamp, λ ≥ 395 nm 8.0 wt% Pt 0.1 M ascorbic acid 3.2 10[thin space (1/6-em)]100 Nat. Chem., 2018, 10, 1180
TpDTz 4,4′-(Thiazolo[5,4-d]thiazole-2,5-diyl)dianiline 300 W Xe lamp, AM 1.5 G Ni-thiolate hexameric cluster 10 vol% TEOA 0.044 941 J. Am. Chem. Soc., 2019, 141, 11082
TpDTP 4,4′′-Diamino-p-terphenyl 300 W Xe lamp, AM 1.5 G Ni-thiolate hexameric cluster 10 vol% TEOA 160 J. Am. Chem. Soc., 2019, 141, 11082
Tp-PDA p-Phenylenediamine 300 W Xe lamp, λ ≥ 420 nm 5.0 wt% Pt Sodium ascorbate 0.76 600 Small, 2021, 17, 2101017
Tp-DBN 2,5-Diaminobenzonitrile 300 W Xe lamp, λ ≥ 420 nm 5.0 wt% Pt Sodium ascorbate 2.12 1800 Small, 2021, 17, 2101017
TpPa-H p-Phenylenediamine (Pa–H) 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 11.13 Chem. Eng. J., 2021, 419, 129984
TpPa-Cl2 2,5-Dichloro-1,4-phenylene-diamine (Pa-Cl2) 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 17 11.73 Chem. Eng. J., 2021, 419, 129984
TpPa-(CH3)2 2,5-Dimethyl-1,4-phenylenedia (Pa-(CH3)2) 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 4.44 Chem. Eng. J., 2021, 419, 129984
TpPa-SO3H 2,5-Diaminobenzenesulphonic acid (Pa-SO3H) 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 3.62 Chem. Eng. J., 2021, 419, 129984
Tp-BD Benzidine 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 7.19 Chem. Eng. J., 2021, 419, 129984
Tp-DTP 4,4′′-Diamino-p-terphenyl 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 4.76 Chem. Eng. J., 2021, 419, 129984
Tp(BT0.05TP0.95)-COF 4,4′-(Benzo-2,1,3-thiadiazole-4,7-diyl)dianiline (BT) and 4,4′-diamino-p-terphenyl (TP) 300 W Xe lamp, λ ≥ 420 nm 5.0 wt% Pt 0.1 M ascorbic acid 2.34 9839 Polym. Chem., 2021, 12, 3250
TAB-TFP-COF 440,400-Boranetriyltris(3,5-dimethylaniline) 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Ascorbic acid 0.69 666.4 J. Mater. Chem. A, 2022, 10, 17691–17698
Tp-2C/BPy2+-COF (19.10%) 2,2′-Bipyridine 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt 0.1 M ascorbic acid 6.93 34[thin space (1/6-em)]600 Angew. Chem. Int. Ed., 2021, 60, 9642–9649
Tta-Tp 4,4′,4′′-(1,3,5-Triazine-2,4,6-triyl) trianiline 300 W Xe lamp, λ > 420 nm 2.2 wt% Pt 0.1 M ascorbic acid 0.64 6286 ChemSusChem, 2022, 15, e202101510
Tapb-Tp 1,3,5-Tris(4-aminophenyl)-benzene 300 W Xe lamp, λ > 420 nm 2.2 wt% Pt 0.1 M ascorbic acid 0.21 1828 ChemSusChem, 2022, 15, e202101510
Ttba-Tp 4,4′,4′′-(1,3,5-Triazine-2,4,6-triyl)-tris(1,1′-biphenyl)trianiline) 300 W Xe lamp, λ > 420 nm 2.2 wt% Pt 0.1 M ascorbic acid 0.44 3386 ChemSusChem, 2022, 15, e202101510
Tta-Tb 4,4′,4′′-(1,3,5-Triazine-2,4,6-triyl) trianiline 300 W Xe lamp, λ > 420 nm 2.2 wt% Pt 0.1 M ascorbic acid 0.04 152 ChemSusChem, 2022, 15, e202101510
Tz-COF-1 2,6-Diaminobenzo [1,2-d:4,5-d]-bisthiazole 300 W Xe lamp, λ > 420 nm 3.0 wt% Pt 0.8 M ascorbic acid 1100 ACS Catal., 2022, 12, 9494–9502
Tz-COF-2 2,6-Diaminobenzo [1,2-d:4,5-d]-bisthiazole 300 W Xe lamp, λ > 420 nm 3.0 wt% Pt 0.8 M ascorbic acid 7100 ACS Catal., 2022, 12, 9494–9502
Tz-COF-3 2,6-Diaminobenzo [1,2-d:4,5-d]-bisthiazole 300 W Xe lamp, λ > 420 nm 3.0 wt% Pt 0.8 M ascorbic acid 6.9 43[thin space (1/6-em)]200 ACS Catal., 2022, 12, 9494–9502
TpPa(Δ)-Cu(II)–COF (S)- or (R)-1-Benzenemethanamine 300 W Xe lamp, λ ≥ 420 nm L-Cysteine 0.78 14[thin space (1/6-em)]720 Nat. Chem., 2022, 13, 5768
Pt@TpBpy-NS 2,2′-Bipyridine-5,5′-diamine 300 W Xe lamp, λ > 420 nm Sodium ascorbate 22[thin space (1/6-em)]666 Nat. Commun., 2023, 14, 593
Pt@TpBpy-2-NS 3,3′-Bipyridine-6,6′-diamine 300 W Xe lamp, λ > 420 nm Sodium ascorbate 9523 Nat. Commun., 2023, 14, 593
Plm-COF2 1H-Phenanthro-[9,10-d] imidazole-5,10-diamine 300 W Xe lamp, λ ≥ 420 nm H2PtCl6 0.1 M ascorbic acid 2.55 7417.5 ACS Appl. Energy Mater., 2023, 6, 1126–1133
Plm-COF1 1H-Phenanthro-[9,10-d] imidazole-5,10-diamine 300 W Xe lamp, λ ≥ 420 nm H2PtCl6 0.1 M ascorbic acid 0.64 528.5 ACS Appl. Energy Mater., 2023, 6, 1126–1133
Tp-DB-H2-COF 4,4′-Diamino-biphenyl benzidine 300 W Xe lamp, λ > 400 nm 3.0 wt% Pt Sodium ascorbate 600 Molecular Catalysis, 2023, 535, 112807
Tp-DB-(CH3)2-COF 3,3′-Dimethyl-benzidine 300 W Xe lamp, λ > 400 nm 3.0 wt% Pt Sodium ascorbate 810 Molecular Catalysis, 2023, 535, 112807
Tp-DB-(OCH3)2-COF 3,3′-Dimethoxy benzidine 300 W Xe lamp, λ > 400 nm 3.0 wt% Pt Sodium ascorbate 1230 Molecular Catalysis, 2023, 535, 112807
Tp-DB-(NO2)2–COF 3,3′-Dinitro-benzidine 300 W Xe lamp, λ > 400 nm 3.0 wt% Pt Sodium ascorbate 15 Molecular Catalysis, 2023, 535, 112807
FOO-COF 2,7-Diamino-9 H-fluoren-9-one 300 W Xe lamp, λ ≥ 420 nm Pt Ascorbic acid 20.5 119[thin space (1/6-em)]100 Appl. Catal., B., 2023, 330, 122581
FO-COF 2,7-Diamino-fluorene 350 W Xe lamp, λ > 400 nm Pt Ascorbic acid 12[thin space (1/6-em)]300 Appl. Catal., B., 2023, 330, 122581
[thin space (1/6-em)]
Bipyridine-based COFs
TPCBP B-COF 1,4-Dibromobutane 300 W Xe lamp, λ ≥ 420 nm TEOA 79.69 1029 ACS Appl. Mater. Interfaces, 2023, 15, 18836–18844
BTT-BPy-PCOF(AC) Benzotrithiophene-2,5,8-tricarbaldehyde 300 W Xe lamp, AM 1.5 G 3.32 wt% Pt 0.1 M ascorbic acid 21[thin space (1/6-em)]200 Angew. Chem. Int. Ed., 2023, 62, e202300224
[thin space (1/6-em)]
Others COFs
ODA-COF 5,5′-(2,5-Diethoxy-1,4-phenylene)-bis(2-phenyl-1,3,4 oxadiazole) 300 W Xe lamp, λ > 420 nm 9.0 wt% Pt TEOA 0.42 2615 Angew. Chem. Int. Ed., 2022, 61, e202115655
H–COF N 1,N4-Di((E)–benzylidene)-2,5-diethoxytere-phthalohydrazide 300 W Xe lamp, λ > 420 nm 9.0 wt% Pt TEOA 609 Angew. Chem. Int. Ed., 2022, 61, e202115655
RuCOF-ETTA Ru(bpy-oEt)3(PF6)2 300 W Xe lamp, λ > 420 nm H2PtCl6 0.01 M ascorbic acid 6.95 6429 Angew. Chem. Int. Ed., 2022, 61, e202208791
RuCOF-TPB Ru(bpy-oEt)3(PF6)2 300 W Xe lamp, λ > 420 nm H2PtCl6 0.01 M ascorbic acid 20[thin space (1/6-em)]308 Angew. Chem. Int. Ed., 2022, 61, e202208791
RuCOF-ETTBA Ru(bpy-oEt)3(PF6)2 300 W Xe lamp, λ > 420 nm H2PtCl6 0.01 M ascorbic acid 14[thin space (1/6-em)]745 Angew. Chem. Int. Ed., 2022, 61, e202208791
Zn@H-TpPa p-Phenylene-diamine 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 28[thin space (1/6-em)]000 ChemCatChem, 2022, 14, e202101800
v-2d-COF-NO1 2,6-Dimethylbenzo-[1,2-b:4,5-b′]-bisoxazole 250 W Xe lamp, λ > 400 nm 8.0 wt% Pt DIPEA 1970 J. Am. Chem. Soc., 2022, 144, 13953–13960
V-2d-COF-NO2 2,6-Dimethylbenzo-[1,2-b:5,4-b′]-bisoxazole 250 W Xe lamp, λ > 400 nm 8.0 wt% Pt DIPEA 863 J. Am. Chem. Soc., 2022, 144, 13953–13960
v-2d-COF-NS1 2,6-Dimethyl-[1,3]thiazolo-[5,4-f][1,3]benzo-thiazole 100 mW cm−2, λ > 420 nm 8.0 wt% TEA 4400 ACS Catal., 2023, 13, 1089–1096
BTT-PDA Benzene-1,4-diamine 150 W Xe lamp, λ > 420 nm 3.0 wt% Pt 0.1 M ascorbic acid 2040 Angew. Chem. Int. Ed., 2023, 62, e202217416
BTT-NDA Naphthalene-2,6-diamine 150 W Xe lamp, λ > 420 nm 3.0 wt% Pt 0.1 M ascorbic acid 5.42 5220 Angew. Chem. Int. Ed., 2023, 62, e202217416
BTT-AnthDA Anthracene-2,6-diamine 150 W Xe lamp, λ > 420 nm 3.0 wt% Pt 0.1 M ascorbic acid 4230 Angew. Chem. Int. Ed., 2023, 62, e202217416
BTT-BPhDA [1,1′-Biphenyl]-4,4′-diamine 150 W Xe lamp, λ > 420 nm 3.0 wt% Pt 0.1 M ascorbic acid 3270 Angew. Chem. Int. Ed., 2023, 62, e202217416


3.2 Choice of synthetic methods for COF preparation

Several synthetic methods have been utilized for preparing crystalline COFs in high-phase purity.86 The solvothermal route is the most commonly employed synthetic technique to prepare crystalline COFs. It involves a condensation reaction between the organic building blocks in a vacuum-sealed tube at a specific temperature under solvent-generated pressure. The first landmark synthesis of COF-1 and COF-5 was achieved by Yaghi and coworkers by self-condensation of 1,4-benzene diboronic acid (BDBA) and its cocondensation with 2,3,6,7,10,11-hexahydroxytriphenylene (HHTP), respectively, under solvothermal conditions.43 Herein, the solvothermal reaction carried out in a sealed tube promotes a condensation reaction between the precursors with controlled dehydration and diffusion of the molecules into the solution, facilitating the nucleation of crystalline material. The use of a closed reaction system sustains the availability of water for maintaining reversible conditions conducive to the growth of COF crystallites. Although the solvothermal method yields highly crystalline and porous COFs, it can be performed only at small scales with harsh reaction conditions for longer reaction times. Furthermore, many new approaches, such as microwave, mechanochemical, and room temperature solution methods, have been developed for COF synthesis.87

In particular, the microwave-assisted solvothermal route was found to be beneficial for the rapid synthesis of porous COFs in good yield and crystallinity. In this regard, Wang and coworkers demonstrated microwave-assisted solvothermal synthesis of crystalline TpPa-COF from 1,3,5-triformylphloroglucinol (Tp) and p-phenylenediamine (Pa) precursors in a mesitylene/1,4-dioxane/acetic acid (3[thin space (1/6-em)]:[thin space (1/6-em)]3[thin space (1/6-em)]:[thin space (1/6-em)]1) solvent mixture at 100 °C within 60 min.88 The ionothermal synthetic route has been employed to rapidly prepare 3D COFs by utilizing ionic liquid (IL) as a green solvent under ambient temperature and pressure conditions.89 In this direction, Fang and coworkers reported the synthesis of 3D-IL-COFs using 1-butyl-3-methylimidazolium bis((trifluoromethyl)sulfonyl)imide ionic liquid, which acts as solvent as well as catalyst.89 The Schiff base reaction of tetrakis(4-formylphenyl)methane (TFPM) with C2 symmetric linear linkers with increasing chain length, p-phenylenediamine (6.1 Å), 4,4′-diaminobiphenyl (10.7 Å) and 4,4′′-diamino-p-terphenyl (15.3 Å) afforded 3D interpenetrated COFs named 3D-IL-COF-1–3 at room temperature within 12 h. The microwave-assisted and ionothermal methods offer an eco-friendly route for rapid and scalable synthesis of COFs over the traditional solvothermal process, which takes several days. However, the general applicability of these methods is lower than that of solvothermal methods.

In addition, room temperature synthesis of an imine-based COF (RT-COF-1) was achieved by a Schiff base reaction between 1,3,5-tris(4-aminophenyl)benzene (TAPB) and 1,3,5-benzenetricarbaldehyde (BTCA) using acetic acid along with m-cresol or DMSO.90 Furthermore, the fabrication of RT-COF on various surfaces to obtain submicron patterns of COF by using ink-jet printing and wet lithography techniques has also been demonstrated. In another instance, Dichtel and coworkers reported rapid low-temperature synthesis of imine-linked COFs named TAPB-PDA by Sc(OTf)2-catalyzed reaction of 1,3,5-tris(4-aminophenyl)benzene (TAPB) with terephtalaldehyde (PDA) in 1,4-dioxane/mesitylene (4[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v) at RT within 30 min.91 This method was also extended to synthesize several heteroatom-containing COFs, demonstrating the general applicability of the synthetic process. To develop a rapid, eco-friendly, and solvent-free synthetic route, Banerjee and coworkers employed a mechanochemical route for synthesizing chemically and thermally stable COFs (TpPa-1/2) by manually grinding the precursors using a mortar and pestle within 45 min.92 This method yielded COFs with moderate crystallinity, marking a promising route for synthesizing COFs under environmentally friendly mild conditions. Banerjee's group further extended this synthetic strategy for the rapid synthesis of highly crystalline and ultra-porous COFs in seconds by utilizing a novel salt-mediated (p-toluenesulfonic acid) crystallization approach.93 In addition, continuous synthesis of COFs was achieved using a twin-screw extruder, and the resulting COFs were processed into different shapes mimicking the ancient terracotta process. Overall, from the aforementioned discussions, it is evident that one can achieve a rapid synthesis of COFs under eco-friendly, mild, solvent-free conditions. These promising reports pave the way for scaling COF synthesis for large-scale production.

However, most of these rapid synthesis methods yield COFs with reduced crystallinity and surface area, leading to lower photocatalytic performance. To summarize, COFs synthesized by the solvothermal route possess high crystallinity and are reported to show enhanced hydrogen evolution performance over their amorphous counterparts (see below).77

3.3 Role of organic moieties as building units

As discussed in the previous sections, COFs are formed by strong covalent bonding between the organic building blocks comprising light elements (B, C, H, N, O, S, etc.) through a reversible condensation reaction. The self-healing of COFs and their thermodynamically controlled covalent bonding results in long-range-ordered crystalline structures.54,75,94 Thus, the network structure and properties of the resulting COFs will depend on the type of linkers used. The various organic linkers utilized in constructing photocatalytic COFs for hydrogen generation are shown in Fig. 6. Most COFs are prepared by the Schiff-base condensation reaction between aromatic amine and aldehyde ligands, both simple and substituted. Most of the linkers employed for COF synthesis have aromatic conjugation. The extent of aromaticity increases from simple phenyl linkers to di- and poly-phenyl-based linkers, further enhancing the visible light absorption properties of the material. In addition, it has been well established in the literature that an increase in the linker's aromatic conjugation results in enhanced COF visible-light photocatalytic activity.95 Thus, highly π-conjugated systems act as strong electron donors and possess a narrow bandgap that promotes the effective separation and migration of photogenerated electrons and holes to the surface under visible light irradiation.56 Consequently, COFs designed from linkers with extended conjugation have shown efficient hydrogen evolution over their lower conjugated counterparts.96
image file: d3ta02189k-f6.tif
Fig. 6 Representative linkers employed for the design of photocatalytic COFs and CTFs for water splitting.

It has been well demonstrated that COFs constructed from a judicious choice of organic linkers consisting of electron-donor and electron-acceptor moieties (donor–acceptor COFs) are beneficial in promoting facile transport of photogenerated electrons with longer lifetimes.97 Consequently, the donor–acceptor COFs formed by amine-based donor-type linkers such as pyrene, porphyrins, triphenyl, etc., and aldehyde-based acceptor-type ligands having heteroatoms or electron-withdrawing substituents (–Cl, –F, –CN) are reported to exhibit enhanced HER activity (Fig. 6).98 Herein, by virtue of donor–acceptor combination, the migration of photogenerated electrons from the donor to the acceptor ligand is favored by a push–pull effect.60

3.4 Linkage chemistry for stable COFs

The chemical stability and durability of the synthesized COFs depend on the nature of the covalent linkage formed between the organic precursors (linkers and knots). It also affects the aromaticity and photostability of the COFs and their photocatalytic performance.99,100 Notably, several covalent linkages, such as boroxine, boronic ester, imine, azine, hydrazine, β-ketoenamine, amide, olefin, dioxin, and ester, are employed for synthesizing COFs (Fig. 7). Among them, boroxine and boronic ester-linked COFs are reported to exhibit lower chemical stability, offering limited applications for photocatalytic water splitting.101 Hence, these COFs have not been utilized for photocatalytic water splitting applications because they are susceptible to water attack and can easily breakdown into individual precursor molecules. Owing to their excellent chemical stability and extended conjugation, imine-, azine-, hydrazone-, β-ketoenamine-, and olefin-linked COFs are preferred for photocatalytic HER applications.36 The various types of linkages employed for the synthesis of COFs and their porosity, chemical stability, photocatalytic performance, and durability are shown in Fig. 7. In addition, covalent triazine frameworks (CTFs) have also been applied for efficient and durable water-splitting performance (see below).42,102
image file: d3ta02189k-f7.tif
Fig. 7 Various linkages, their properties and representative examples of the COFs employed for the synthesis of photocatalytically active COFs.

3.5 Functionalization of COFs for enhanced photocatalytic hydrogen generation

To enhance photocatalytic HER performance, pristine COFs are often functionalized by immobilizing catalytically active metal ions, complexes, or semiconductor nanomaterials (Fig. 8). In this context, COFs functionalized with -azo, -azine, and bipyridine are suitable for anchoring catalytically active species. For instance, bipyridine-functionalized COFs are incorporated with reactive metal ions by postsynthetic modification (PSM) to obtain hybrid COFs with enhanced hydrogen generation activity.79,103,104 A similar observation of enhanced photocatalytic water oxidation activity has been reported for a cobalt-embedded bipyridine-based COF under visible light irradiation.105 In another example, a β-ketoenamine-linked COF composed of a 2,2′-bipyridine linker was quaternized with dibromo alkanes into cyclic diquats, and its photocatalytic activity was studied. The resultant COF showed outstanding hydrogen generation activity (rate of 34[thin space (1/6-em)]600 μmol h−1 g−1) compared to the parent COF.106 In addition, bipyridine-functionalized COFs have been utilized for covalent anchoring of noble metal (Ru/Rh) complexes for photocatalytic applications. In addition, anchoring of a cadmium sulfide (CdS) semiconductor photocatalyst in TpPa COF to obtain a CdS/TpPa hybrid with enhanced hydrogen evolution activity over the parent COF and CdS has been demonstrated.107 The crystalline, rigid network of COFs supports the formation of highly exposed catalytic sites in the large 1D channels – facilitating facile diffusion of the reactants and products.
image file: d3ta02189k-f8.tif
Fig. 8 Various strategies employed for the design of COF-based hybrid photocatalysts for photocatalytic hydrogen evolution.

On the other hand, the integration of COFs with carbon-based 2D materials such as graphene, graphene oxide (GO), and graphitic carbon nitride (g-C3N4) has been found to be beneficial in enhancing the photocatalytic performance of COFs (Table 3)108,109 The layered structure of both COFs and carbon materials facilitates facile charge transport by suppressing the charge recombination associated with individual materials, leading to enhanced hydrogen generation efficiency. In this regard, an imine-linked COF formed from 4,4′,4′′-(1,3,5-triazine-2,4,6-triyl)trianiline (TTA) and 1,3,5-triformylphloroglucinol (TP) was functionalized in situ with g-C3N4 (CN) to obtain a CN-COF hybrid. The hybrid obtained showed excellent photocatalytic hydrogen generation activity with a steady rate of 10.1 mmol g−1 h−1 under visible light irradiation (425 nm).110 The enhanced HER performance of the hybrid has been attributed to broader absorption of visible light and facile separation/migration of photogenerated electron–hole pairs aided by the heterojunction formation between COF and g-C3N4 with optimal band edge positions.

Table 3 The hydrogen evolution properties of COF-hybrid materials
Name of COF-hybrid photocatalyst Name of COF Dopant used Light intensity Conc. of cocatalyst Sacrificial donor AQY (%) Hydrogen evolution rate (μmol h−1 g−1) Reference
CdS-COF COF(TpPa-2) CdS 400 W Xe lamp, λ ≥ 420 nm 0.5 wt% Pt 1 mL lactic acid 4.2 3678 Chem.–Eur. J., 2014, 20, 15961–15965
N2-COF/Co1 N2-COF Chloro(pyridine)-cobaloxime (Co1) λ ≥ 420 nm 8.0 wt% Pt 1 μL TEOA 782 J. Am. Chem. Soc., 2017, 139, 16228
NH2-UiO-66/TpPa-1-COF TpPa-1-COF NH2-UiO-66 MOF λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 23[thin space (1/6-em)]410 Angew. Chem. Int. Ed., 2018, 130, 12282
NH2-UiO-66/TAPT-TP-COF TAPT-TP-COF NH2-UiO-66 MOF AM 1.5 G 3.0 wt% Pt 10% TEOA 5.7 8440 ACS Appl. Mater. Interfaces, 2021, 3, 29916
MOF-808@TpPa-1-COF TpPa-1-COF NH2-UiO-808 MOF 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 10.23 11[thin space (1/6-em)]880 J. Mater. Chem. A, 2021, 9, 16743
MoS2-TpPa-1-COF TpPa-1-COF MoS2 300 W Xe lamp, λ ≥ 420 nm Ascorbic acid 0.76 5585 J. Mater. Chem. A, 2019, 7, 20193
Ni(OH)2/COF-TpPa-2 COF-TpPa-2 Ni(OH)2 300 W Xe lamp, λ ≥ 420 nm Sodium ascorbate 1890 Chem. Eng. J., 2020, 379, 122342
α-Fe2O3/TpPa-2-COF TpPa-2-COF α-Fe2O3 300 W Xe lamp, λ ≥ 420 nm Sodium ascorbate 0.137 3770 J. Mater. Chem. A, 2020, 8, 4334
TiO2-TpPa-1-COF TpPa-1-COF TiO2 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 7.6 11[thin space (1/6-em)]190 Appl. Catal., B, 2020, 266, 118586
COF-CN TpBD g-C3N4 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 15.09 12[thin space (1/6-em)]800 ACS Appl. Mater. Interfaces, 2020, 12, 51555
rGO-TpPa-1-COF TpPa-1-COF rGO (reduced graphene oxide) 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 7.53 11[thin space (1/6-em)]980 J. Mater. Chem. A, 2020, 8, 8949
TBTA/g-C3N4 TBTA g-C3N4 300 W Xe lamp, λ ≥ 420 nm 2.08 wt% Pt Ascorbic acid 26[thin space (1/6-em)]040 Catal. Sci. Technol., 2021, 11, 2616
PEG@BT-COF BT-COF PEG (polyethylene glycol) 300 W Xe lamp, λ ≥ 420 nm Ascorbic acid 11.2 11[thin space (1/6-em)]140 Nature Comm., 2021, 12, 3934
PdTCPP⊂PCN415(NH2)/TpPa (composite 2) TpPa-COF PdTCPP⊂PCN415(NH2) 300 W Xe lamp, λ ≥ 420 nm 1.2 wt% Pt Ascorbic acid 5.9 13[thin space (1/6-em)]980 Angew. Chem. Int. Ed., 2022, 61, e202114071
CN/TMP TMP COF g-C3N4 300 W Xe lamp, λ ≥ 420 nm Pt TEOA 4.07 2057 Appl. Catal., B, 2022, 315, 121568
TiO2-x/TpPa-1-COF (6[thin space (1/6-em)]:[thin space (1/6-em)]4) TpPa-1-COF TiO2 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Sodium ascorbate 6.7 15[thin space (1/6-em)]330 Chem. Eng. J., 2022, 446, 137213
Pt/rGO(20%)/TpPa-1-COF TpPa-1-COF diamine rGO (reduced graphene oxide) 300 W Xe lamp, λ > 420 nm 3.0 wt% Pt Sodium ascorbate 13.82 19[thin space (1/6-em)]590 J. Colloid interface Sci., 2022, 608, 2613–2622
MS-c@TpPa-1 TpPa-1-COF [Mo3S13]2− nanoclusters 300 W Xe lamp, λ > 420 nm [Mo3S13]2− nanoclusters Sodium ascorbate 0.54 528 Chem. Eng. J, 2022, 446, 136883
TpPa-1-COF/ZnIn2S4 TpPa-1-COF ZnIn2S4 300 W Xe lamp, λ ≥ 420 nm 0.35 M Na2S + 0.25 M Na2SO3 2.08 853 ACS Appl. Energy Mater., 2023, 6, 1103–1115
Ni-COF-SCAU-1 COF-SCAU-1 Ni(acac)2 300 W Xe lamp, λ ≥ 420 nm 3.0 wt% Pt Ascorbic acid 43.2 197[thin space (1/6-em)]460 Adv. Energy Mater., 2023, 13, 2203695


Recently, Yan and coworkers demonstrated the hybridization of TpPa-1-COF with a metal–organic framework (NH2-UiO-66) to obtain an NH2-UiO-66/TpPa-1-COF hybrid with improved stability and photocatalytic activity.111 The hybrid was fabricated via the Schiff base reaction of the COF precursors 4,4′,4′′-(1,3,5-triazine-2,4,6-triyl)trianiline (TAPT) and 2,4,6-triformylphloroglucinol (TP) in the presence of NH2-UiO-66. Interestingly, the hybrid material exhibited improved stability in an alkaline environment and a high photocatalytic hydrogen generation rate (8.44 mmol h−1 g−1) under visible light (400 nm). Notably, the hybrid's hydrogen generation activity was higher than that of pure MOF and COF materials (see below). Hence, hybridizing COFs with appropriate materials has proven to be a promising strategy for generating hybrid materials with improved chemical/photostability, light-harvesting ability, and photocatalytic activity for sustainable hydrogen generation.

4. Photocatalytic hydrogen generation activity of COFs

The following section discusses the major types of COFs known for the photocatalytic hydrogen evolution reaction (HER) from water.

4.1 Triazine-based COFs and CTFs

In addition to triazine-containing COFs, covalent triazine frameworks (CTFs) have been vastly used for water splitting, which is discussed in the following sections.
4.1.1 Triazine-based COFs. Notably, COFs composed of pyrazine, triazine, tetrazine, pyrimidine, and heptazine-based nitrogen-rich linkers exhibit modest HOMO–LUMO separation, resulting in improved visible light harvesting and HER performance. Theoretical studies have shown that photoexcited triazine molecules can abstract a hydrogen atom from water via a hole transfer mechanism involving triazinyl–hydroxyl biradical formation. Thus, in a triazine system, the hydrogen-bonded water molecule decomposes into hydrogen and hydroxyl radicals in a biphotonic photochemical reaction facilitating water splitting.

To take advantage of these properties of triazine, Lotsch and coworkers demonstrated the application of a triazine COF for photocatalytic water hydrogen generation.77 The TFPT-COF was designed by using 1,3,5-tris-(4-formylphenyl)triazine (TFPT) and 2,5-diethoxy-terephthalohydrazide (DETH) building blocks (Fig. 9a). The coupling of the triazine precursor with DETH resulted in a hydrazone-linked COF featuring good chemical stability, crystallinity and visible light harvesting properties (band gap of 2.8 eV). The TFPT-COF showed a Brunauer–Emmett–Teller (BET) surface area of 1603 m2 g−1 and pore volume of 1.03 cm3 g−1. Indeed, TFPT-COF exhibited good HER activity with a moderate hydrogen evolution rate of 230 μmol h−1 g−1 with Pt as the cocatalyst and sodium ascorbate as the SED (Fig. 9b). Interestingly, the use of triethanolamine as SED led to an increase in the HER rate to 1970 μmol h−1 g−1 (Fig. 9b). Further studies on the effect of incident light on the HER revealed the best H2 generation activity at 400 nm light with the highest apparent quantum efficiency (Fig. 9c). The recyclability studies showed partial loss of crystallinity of COF after using it for 95 h due to exfoliation of COF in water.


image file: d3ta02189k-f9.tif
Fig. 9 Synthesis and photocatalytic applications of TFPT-COF. (a) Scheme of the synthesis of TFPT-COF. (b) Hydrogen evolution performance of TFPT-COF using sodium ascorbate and triethanolamine as sacrificial electron donors. (c) Wavelength-specific hydrogen generation of Pt-immobilized TFPT-COF. Reproduced with permission from ref. 77 Copyright© 2014, Royal Society of Chemistry.

In subsequent pioneering efforts to explore the structure–property relationship, Lotsch and coworkers studied the effect of tailoring the structure of COF at the molecular level by varying the number of nitrogen atoms in the precursor on the hydrogen evolution performance.78 They prepared a series of NX-COFs (X = 0–3) by condensation reaction of hydrazine with triphenylaryl aldehydes by introducing N atoms in the central phenyl ring, N = 0 (phenyl), N = 1 (pyridyl), N = 2 (pyrimidyl) and N = 3 (triazine) (Fig. 10a). The substitution of the C–H moiety by N resulted in loss of planarity of the molecule due to a change in the dihedral angle between the central aryl ring and peripheral phenyl rings. The COFs exhibited high crystallinity (Fig. 10b) and were applied for photocatalytic water splitting. All the NX-COFs showed photocatalytic water splitting activity with hydrogen generation rates of 23, 90, 438, and 1703 μmol h−1 g−1 for N0, N1, N2, and N3-COFs, respectively. Interestingly, the hydrogen evolution rate increases with an increase in the number of nitrogen substitutions with a fourfold increase per nitrogen atom addition, and N3-COF showed the best activity (Fig. 10c). The increase in the nitrogen content led to lowering of the HOMO energy, facilitating effective separation of photoinduced charge carriers and higher hydrogen evolution performance.


image file: d3ta02189k-f10.tif
Fig. 10 Synthesis and applications of Nx-COFs for water splitting. (a) Scheme of the synthesis of Nx-COFs from Nx-aldehydes and hydrazine. PXRD patterns (b) and hydrogen evolution performance (c) of Nx-COFs. Reproduced with permission from ref. 78 Copyright© 2015, Springer Nature.

In another significant contribution, Lotsch and coworkers reported the application of a chloro(pyridine)cobaloxime (Co1) molecular complex as a cocatalyst to boost the photocatalytic HER performance of an azine-linked N2-COF, replacing the traditional Pt-based cocatalyst.79 Among the various metal complex-based cocatalysts known, cobaloximes are most efficient, offer a low overpotential for hydrogen generation and can be covalently anchored with COFs. The photocatalytic water splitting study of N2-COF in the presence of cobaloxime cocatalyst showed a hydrogen evolution rate of 160 μmol h−1 g−1. Since the complex is unstable and undergoes ligand exchange, adding a dimethylglyoxime (dmgH2) ligand during catalysis enhanced the hydrogen evolution activity. Consequently, the hydrogen evolution rate of N2-COF was enhanced to 782 μmol h−1 g−1 with the addition of 4.69 mM dmgH2 during the water splitting reaction, demonstrating noble metal (Pt)-free cocatalyst-supported COF-based H2 generation. Lotsch's group extended the concept for immobilizing a chloro(pyridine)cobaloxime complex in hydrazone-based COF-42 via PSM to achieve improved and sustained photocatalytic activity.112 The covalent anchoring of the complex cocatalyst was facilitated by coordination to the pyridine group in the COF, resulting in hybrid COFs (Co1a and Co1b). The photocatalytic HER studies of the hybrids revealed enhanced hydrogen evolution compared to pristine COF. The higher HER performance of the hybrid has been attributed to the close contact between the COF and cobaloxime complex in the pore walls of the COF, facilitating facile charge transfer between the complex cocatalyst and COF.

In conjugated polymeric materials, donor–acceptor systems show efficient light absorption and charge transport, essential for an effective photocatalytic water-splitting application. Utilizing this strategy, Dong and coworkers designed two benzothiadiazole-based donor–acceptor COFs, BT-TAPT-COF and BDF-TAPT-COF, featuring good crystallinity, chemical stability, and light harvesting properties.82,113 The COFs were prepared by imine condensation of tris-(4-aminophenyl)triazine (TAPT) with two different aldehydes, 4,4′-(benzothiadiazole-4,7-diyl)dibenzaldehyde (BT-CHO) and 4,4′-(benzo[1,2-b:4,5-b′]difuran-4,8-diyl)dibenzaldehyde (BDF-CHO). In BDF-TAPT-COF, the triazine molecule acts as an electron acceptor, and the benzothiadiazole or benzodifurane molecule acts as an electron donor (Fig. 11a). The water-splitting experiments revealed higher hydrogen generation by BDF-TAPT-COF (1390 μmol h−1 g−1) than by BT-TAPT-COF (949 μmol h−1 g−1) (Fig. 11c). Additionally, UV-Vis diffuse reflectance spectrum also supports the enhanced light-harvesting property of BDF-TAPT-COF (Fig. 11b). This has been attributed to the higher electron-donating ability of the benzothiadiazole moiety rendering efficient charge separation and transfer for the reduction of water to hydrogen evolution.


image file: d3ta02189k-f11.tif
Fig. 11 Synthesis and applications of BDF-TAPT-COF for photocatalytic hydrogen generation. (a) Synthesis of BDF-TAPT-COF following the solvothermal method. (b) UV-Vis spectra of BDF-TAPT-COF, TAPT and BDF-CHO linkers. Inset: Tauc plots for the determination of the optical band gaps. (c) Hydrogen evolution studies using BDF-TAPT COF. Reproduced with permission from ref. 113 Copyright 2021, Royal Society of Chemistry.

In a similar concept, Yan and coworkers reported the construction of fluorine-functionalized TFA-COF by the Schiff-base condensation reaction of 2,3,5,6-tetrafluoroterephthaldehyde (TFA-CHO) and 2,4,6-tris(4-aminophenyl)-1,3,5-triazine.114 The COF showed a high surface area (2490 m2 g−1) and optimum band gap (2.40–2.82 eV) owing to the electron-withdrawing effect of the fluorinated aldehyde. Interestingly, the TFA-COF showed HER with a rate of 80 μmol g−1 h−1, which is higher than that of nonfluorinated COF (TA-COF) having a terephthaldehyde (TA) linker with a H2 generation rate of 10 μmol g−1 h−1. The higher catalytic activity of TFA-COF has been ascribed to a reduced band gap resulting in enhanced charge separation and migration aided by electron-accepting fluorine atoms. This work highlights the importance of the choice of linkers in tuning the photocatalytic performance of COFs.

Recently, Thomas and coworkers demonstrated the effect of protonation as a promising means to improve the photocatalytic hydrogen generation performance of an imine-based COF.83 They prepared three donor–acceptor type imine-linked COFs named TtaTfa, TpaTfa, and TtaTpa from 2,4,6-tris(4-aminophenyl)triazine (Tta), tris(4-formylphenyl)amine (Tfa), 1,3,5-tris(4-formylphenyl)benzene (Tpa-CHO) and 1,3,5-tris(4-aminophenyl)benzene (Tpa-NH2) building blocks (Fig. 12a). The COFs exhibited high crystallinity and a porous structure with well-defined 2D hexagonal honeycomb-like pore channels. The protonated COFs obtained by treatment with ascorbic acid showed enhanced hydrogen generation activity under visible light (>420 nm). For instance, the protonated COF, TtaTfa, showed the best hydrogen evolution activity with a rate of 20.7 mmol g−1 h−1. However, TpaTfa and TtaTpa showed hydrogen generation rates of 14.9 and 10.8 mmol g−1 h−1, respectively (Fig. 12b). Interestingly, the HER performance of the COFs was negligible when TEOA was used as a SED, highlighting the role of protonation on the hydrogen evolution activity (Fig. 12c). Here, protonation of COF led to broader absorption of light and better charge migration aided by planarity of the molecule, resulting in higher hydrogen generation activity over the nonprotonated counterpart (Fig. 12d).


image file: d3ta02189k-f12.tif
Fig. 12 (a) Synthesis scheme for protonated 2D imine-linked COFs. (b and c) Comparison of the HER performance of protonated COFs carried out using AC and TEOA as SED and (d) mechanism of protonation of pristine COFs in L-AC. Reproduced with permission from ref. 83 Copyright© 2021, Wiley-VCH.

Donor–acceptor (D–A) type porous and conjugated materials are widely used in water splitting applications as the optoelectronic properties of donor–acceptor materials can be easily modulated by controlling the intramolecular charge transfer (ICT) from the electron-donating unit (D) to the electron-accepting unit (A). In this regard, Liu and co-workers have demonstrated the structural advantages of three-component donor–π–acceptor (TCDA) COFs, where three-component COFs have shown higher activity compared to analogous two-component COFs consisting of a node and linker, respectively.115 Following the multi-component strategy for the synthesis of two isomorphic TCDA COFs, the sp2-carbon-linked COF-JLU35 was synthesised by Knoevenagel condensation between electron-rich benzo[1,2-b:3,4-b′:5,6-b′′]trithiophene-2,5,8-tricarbaldehyde (BTT), electron-poor 1,3,5-tris-(4-formyl-phenyl)triazine (TFPT) and 1,4-phenyldiacetonitrile (PDAN) in the presence of a catalytic amount of Cs2CO3. On the other hand, imine-linked COF-JLU36 was prepared by Schiff base condensation of benzo[1,2-b:3,4-b′:5,6-b′′]trithiophene-2,5,8-tricarbaldehyde (BTT), 1,3,5-tris-(4-formylphenyl)triazine (TFPT) and 1,4-phenylenediamine (PDA) in the presence of acetic acid catalyst (Fig. 13a). A sp2-carbon-linked COF-JLU35 (λ = 616 nm, τ = 2.68 ns) exhibited a red-shifted absorption spectrum and longer excited state lifetime compared to COF-JLU36 (λ = 548 nm, τ = 0.73 ns), indicating enhanced π-electron efficiency and improved charge separation in the fully conjugated sp2-carbon bond (Fig. 13b and d). Photocatalytic hydrogen evolution was therefore carried out under visible light irradiation (Fig. 13c). Surprisingly, COF-JLU35 exhibited an impressive HER of up to 70[thin space (1/6-em)]800 μmol h−1 g−1 compared to COF-JLU36 (23[thin space (1/6-em)]600 μmol h−1 g−1) (Fig. 13e). This increase in activity can be attributed to the exceptional charge separation, broader light absorption and enhanced photocatalytic activity exhibited by the tri-component donor–π–acceptor COFs.


image file: d3ta02189k-f13.tif
Fig. 13 Synthesis and applications of multicomponent COF-JLU35 and COF-JLU36 for photocatalytic hydrogen generation. (a) Synthesis of COF-JLU35 and COF-JLU36 following the solvothermal method. (b) UV-Vis spectra of COF-JLU35 and COF-JLU36. (c) Study of AQY for different specific wavelengths using COF-JLU35. (d) Photoluminescence decay plot of COF-JLU35 (red) and COF-JLU36 (blue). (e) Hydrogen evolution performances of TCDA COFs. Reproduced with permission from ref. 115 Copyright© 2023, American Chemical Society.
4.1.2 Triazine-based CTFs. Covalent triazine frameworks (CTFs) are a subclass of COFs and have gained special interest owing to their advantages, such as high nitrogen content, chemical tunability, high surface area, permanent microporosity and thermal/chemical stability. The CTFs mainly comprise triazine moieties linked together to form conjugated structures. They are synthesized by acid-catalyzed trimerization of aromatic nitriles or condensation reaction of benzylamine, aromatic aldehydes, or amides. Owing to the extended conjugation and light absorption in the visible region, the application of CTFs for hydrogen generation from water has been studied. The following section discusses selected examples of CTFs known for HER activity.

Early studies on the application of CTF-1 as a photocatalyst for hydrogen evolution were conducted by Wu and coworkers.116 The trifluoromethanesulfonic acid (TFMSA)-catalyzed trimerization of 1,4-dicyanobenzene at RT resulted in CTF-1. It showed absorption at 350–420 nm with an optical band gap of 2.94 eV. The photocatalytic studies showed a hydrogen evolution rate of 200 μmol h−1 g−1 under visible light irradiation (λ ≥ 420 nm) in the presence of Pt as a cocatalyst and triethanolamine as the SED.

In a major development, Thomas and coworkers studied the effect of shortening the reaction time from 40 h to as little as 30 min on the photocatalytic hydrogen evolution performance of CTF-1_X (X = 2.5, 5, 10, 15, 20, and 30 min) by carrying out the synthesis in molten ZnCl2 at 400 °C (Fig. 14a).117 It was found that decreasing the reaction time led to reduced carbonization of the pre-CTF, rendering CTF-1 with improved crystallinity and catalytic activity. Indeed, CTF-1 prepared with just 10 min of reaction time showed impressive hydrogen generation activity with a rate of 1072 μmol h−1 g−1 (AQY of 9.2%) under visible light (>420 nm) in the presence of the Pt cocatalyst and TEOA as the sacrificial agent (Fig. 14b and c). The photocatalytic activity was found to be consistent over a period of 80 hours, proving the stability of the catalyst under water-splitting reaction conditions.


image file: d3ta02189k-f14.tif
Fig. 14 Synthesis and applications of CTFs for water splitting. (a) Two-step synthesis of CTF in short reaction times. (b) Photocatalytic hydrogen evolution performance of pre-CTF and CTF-1_X min materials under visible light irradiation. (c) Comparison of the photocatalytic hydrogen evolution performance of the pre-CTF and CTF-1_X min catalysts under visible light irradiation. Reproduced with permission from ref. 117 Copyright 2017, Royal Society of Chemistry.

In another important demonstration of the effect of functionalization on the water-splitting activity of CTFs, Wu and coworkers synthesized phosphorus (P)-incorporated CTF by thermal treatment of CTF-1 with red phosphorus.118 The doping of phosphorus atoms in the CTF framework was found to alter the electronic and optical properties of CTF-1, promoting facile separation and migration of photoinduced electron–hole pairs, thereby enhancing the HER performance, which was found to be approximately 4.5 times higher than that of pure CTF. In similar attempts, Li and coworkers developed CTF-based heterostructures (CTF-BT/Th-x) by incorporating electron-withdrawing benzothiadiazole (BT) and electron-donating thiophene (Th) moieties in CTF to accomplish enhanced photocatalytic hydrogen evolution activity.119 The heterostructures were prepared by a sequential polymerization reaction of nitrile precursors catalyzed by trifluoromethanesulfonic acid (TfOH) (Fig. 15a). The integration of the electron-withdrawing (BT) and electron-donating (Th) groups was found to alter the HOMO and LUMO positions in CTF-BT/Th-x (where x represents the benzothiadiazole to thiophene molar ratio) in comparison. The resulting band structure at the heterojunction facilitated facile migration of photogenerated electrons and holes, leading to enhanced photocatalytic hydrogen evolution with a rate of 6600 μmol h−1 g−1 under visible light (420 nm) in the presence of triethanolamine (TEOA) as the sacrificial electron donor and Pt (3 wt%) nanoparticles as cocatalysts (Fig. 15b and c). These reports demonstrate the importance of doping heteroatoms in the CTF backbone to achieve efficient photocatalytic hydrogen generation.


image file: d3ta02189k-f15.tif
Fig. 15 (a) Sequential polymerization strategy for the synthesis of covalently connected CTF-BT/Th, (b) illustration of the band alignment at the heterostructure, and (c) time-dependent photocatalytic hydrogen generation. Reproduced with permission from ref. 119 Copyright© 2019, Wiley-VCH.

For the preparation of highly active COF-based photocatalysts for solar energy harvesting and conversion, the donor–acceptor (D–A) strategy is an effective synthetic approach. Following a similar strategy, Jin and coworkers reported the construction of a series of D–A1–A2 system-based CTFs for photocatalytic hydrogen evolution.120 The frameworks were prepared through copolymerization of 4,7-bis(4-formylphenyl)-2,1,3-benzothiadiazole (M-BT) and 3,6-dicarbaldehyde-N-ethylcarbazole (M-CBZ) in designated ratios with terephthalimidamide dihydrochloride. Thus, a series of ter-CTF-x (x = percentage of M-CBZ unit in CTF system varying from 10 to 90%) were prepared. Of these materials, ter-CTF-0.7 showed the best photocatalytic performance with a hydrogen evolution rate of 966 μmol h−1 (AQY of 22.8%) at 420 nm. Due to the suppressed recombination of the electrons and holes, more delocalized structure for charge migration, and highest charge-transfer/separation efficiency, ter-CTF-0.7 showed efficient hydrogen evolution performance, suggesting the critical role of the organization mode and the distance of the D and A units. Similarly, using a series of bipyridine-based CTFs synthesized through polycondensation, Wisser and coworkers demonstrated a direct correlation of the bipyridine content in CTF with the surface area, band gap, charge separation and surface wettability with water and subsequently on the photocatalytic hydrogen evolution performance.121 Among the series of bipyridine-based materials, CTF-02-Bpy0.66 showed the best performance, reaching the highest hydrogen evolution rate of 7.24 mmol h−1 g−1 under visible light irradiation in the presence of a cocatalyst and a sacrificial electron donor due to the optimum balance between optoelectronic properties and highest hydrophilicity. This report exemplifies that the balance between optoelectronic and surface properties and hydrophilicity determines the photocatalytic activity.

In another major development, Cooper and coworkers demonstrated the synthesis of a structurally diverse family of 39 CTFs by using the Suzuki–Miyaura coupling reaction of 2,4,6-tris[4-(4,4,5,5-tetramethyl-1,3,2-dioxaborolan-2-yl)phenyl]-1,3,5-triazine with various dibromo linkers,122 which includes five-membered aromatic heterocycles, substituted phenylenes, bipyridyl heterofluorenes, azoles, quinoxalines, etc. The photocatalytic hydrogen generation performance of these COFs was correlated to their ionization potential, electron affinity, optical gap, and transmittance. Out of all CTF-xx samples (where xx represents the number of dibromo monomers utilized during the synthesis), CTF-15 was found to show the best hydrogen evolution rate of 2946 μmol h−1 g−1 under visible light (>420 nm) irradiation over 50 hours. The electron affinity, ionization potential, optical band gap, and dispersibility of the CTF particles in solution are essential factors in determining the photocatalytic activity of the materials. These results demonstrated a powerful method to discover new photocatalysts and provided valuable insights into the structure–property relationships across a diverse range of chemical functionalities in the material backbones.

The presence of aromatic triazine linkages and fully conjugated structures enriched with aromatic nitrogen atoms make CTFs appealing in photocatalysis applications. Although many synthetic methods have been reported, most of the CTFs synthesized from those methods are amorphous. To develop crystalline, porous and hydrophilic CTFs using a wide array of building blocks, a polycondensation approach was employed for the construction of CTFs under mild conditions. Typically, the condensation reaction of an aldehyde and an amidine dihydrochloride involving Schiff base formation followed by Michael addition was employed for the construction of CTF-HUSTs in the presence of cesium carbonate in a dimethylsulfoxide-water solvent mixture heated at 60 °C for 12 h, then heated at 80 °C, and then 100 °C for 12 h, separately, before being heated at 120 °C for 3 days.123 The crystalline and porous CTF-HUST-1 showed hydrogen evolution with a rate as high as 1460 μmol h−1 g−1, which is higher than that of CTF-1 synthesized by an ionothermal method (1072 μmol h−1 g−1). Similarly, using in situ formation of aldehyde monomers through the controlled oxidation of alcohols to aldehydes, Tan and coworkers achieved the synthesis of CTF-HUST-C1 by reticulating them with terephthalamidine – demonstrating that the improvement in crystallization lies in decreasing nucleation rates and generating lower concentrations of nuclei.124 Using CTF-HUST-C1, which has much-improved crystallinity, higher thermal stability and layered structures, highly efficient photocatalytic hydrogen evolution performance (5100 μmol h−1 g−1) than the low crystalline CTF-HUST-1 (1460 μmol h−1 g−1) or amorphous CTFs (1072 μmol h−1 g−1) was achieved due to the easier electronic transport and better light absorption.

Furthermore, Jin and coworkers synthesized CTF-HUST-A1 by utilizing a benzylamine-functionalized monomer that initially formed an imine bond, which was further transformed into a triazine moiety.125 Herein, the basicity of the base reagent employed was found to be very crucial in the generation of crystalline COF. CTF-HUST-A1-tBuOK showed superior hydrogen evolution activity with a rate of 9200 μmol h−1 g−1 and AQE of 7.4% under visible light (>420 nm) irradiation along with good stability and recyclability. Tan and coworkers demonstrated the effect of noble metal cocatalysts on hydrogen generation activity by using two isostructural highly crystalline CTFs, named CTF-HC2 and CTF-HC6, with different pyridinic nitrogen contents.126 The CTFs were obtained by following a controlled feeding rate method127 in which the aldehyde solution [4,4′-biphenyl dicarboxaldehyde (2,2′-bipyridine)-5,5′-dicarbaldehyde] was slowly dropped into the amidine (terephthalamidine dihydrochloride) solution. The CTFs were treated with Pd(OAc)2 and H2PtCl6 to generate Pd@CTFs and Pt@CTFs, respectively. Interestingly, the Pd@CTFs were found to exhibit higher HER performance than the Pt@CTFs, which has been attributed to the smaller loading of Pd nanoparticles facilitating better separation of photogenerated electron–hole pairs in comparison to the larger Pt nanoparticle-loaded sample. Hydrogen generation was found to be durable and recyclable for multiple cycles.

Most photocatalytic water splitting studies using COFs are usually performed using materials in the form of suspended powders. Owing to their convenient separation and recycling potential, preparations of COF and CTF films are desired. However, existing synthetic approaches for COF and CTF preparation mainly result in insoluble and unprocessable powders, making their water-splitting applications challenging. To address these issues, Tan and coworkers recently reported an aliphatic amine-assisted interfacial polymerization technique to obtain free-standing and semicrystalline CTF films (Fig. 16a).128 In particular, the dimethylsulfoxide soluble aldehyde monomer was transformed into a dimethylsulfoxide-insoluble imine precursor by reacting with n-hexylamine (Fig. 16b). The initial arrangement of aldehyde monomers is guided via the imine precursor distributed at the DMSO surface, generating a DMSO/air interface at the same time, which is helpful for the preparation of CTF thin films. Using this technique, CTF films with sizes as high as 250 cm2 and average thicknesses up to 30 to 500 nm were prepared (Fig. 16c and d). These films immobilized on a glass support were used for photocatalytic hydrogen generation, which exhibited good photocatalytic hydrogen evolution performance (5.4 mmol h−1 m−2) in the presence of Pt cocatalysts (Fig. 16e–h). The considerably high hydrogen evolution performance was ascribed to the good light absorption, crystalline structure, and large lateral sizes of the films.


image file: d3ta02189k-f16.tif
Fig. 16 Synthesis and applications of CTF films for water splitting. (a) Scheme of the synthesis of CTF films. (b) The fabrication of CTF films on a dimethylsulfoxide surface assisted by an imine precursor. (c) Transparent and free-standing CTF films. (d) Solvent-immersed CTF films. (e) Light absorption spectrum of CTF films. (f) The photocurrent curve and (g) photocatalytic hydrogen evolution performance of CTF films on glass for four cycles. (h) An image of CTF film deposited on a glass plate for photocatalytic hydrogen evolution. Reproduced with permission from ref. 128 Copyright© 2021, Springer Nature.
4.1.3 Heptazine-based COFs. Like triazine moieties, heptazine-based monomers have also been used to synthesize porous frameworks due to their nitrogen-rich, rigid, tricyclic aromatic backbone and their potential photocatalytic applications.129–133 In particular, in 2013, Thomas and coworkers demonstrated the use of heptazine-based microporous polymer networks (HMP-1) as photocatalysts for visible light-driven hydrogen evolution.134 Through a polycondensation reaction of cyanuric chloride with phenylenediamine, the synthesis of heptazine-based polymer (HMP-1) was achieved. Interestingly, HMP-1 exhibited good absorption in the visible region (400–600 nm) due to its aromatic network, which confers photocatalytic properties. The photocatalytic hydrogen generation experiments performed using HMP-1, Pt (3 wt%) as cocatalyst and TEA as SED showed H2 generation at a rate of 2 μmol h−1 at λ > 420 nm. Later, in 2016, the same group reported a donor–acceptor-based heptazine-based polymer, named HMP-3, by replacing phenylenediamine with benzothiadiazole moieties in the network.135 Indeed, the molar ratio of heptazine to benzothiadiazole was varied to obtain HMP-3_4:3 and HMP-3_2:3, which showed relatively higher hydrogen evolution rates of 32 and 31 μmol h−1, respectively. The higher HER of HMP-3 was attributed to the stabilization of photogenerated electrons and holes by a donor–acceptor effect. Motivated by these studies, Tan and coworkers synthesized a series of 2D covalent heptazine-based frameworks (CHFs) by covalently linking heptazine with electron-rich alkynyl, phenyl or alkynyl–phenyl units.136 Furthermore, the overall water splitting performance was investigated using rationally constructed 2D covalent heptazine-based frameworks, which revealed that the 2D CHFs possess tunable band gaps in the range of 2.27–3.26 eV. The CHFs prepared using p-phenyl, p-phenylenediynyl, m-phenylenediynyl and phenyltriynyl showed considerable performance for total water splitting to produce H2 and O2. These studies demonstrate the potential utility of heptazine-based frameworks for photocatalytic water splitting to H2 evolution.137,138

4.2 Pyrene-based COFs

Pyrene-based COFs have attracted significant interest as promising candidates for photocatalytic hydrogen generation from water owing to their extended π-conjugation and visible light absorption properties. In this context, Lotsch and coworkers designed pyrene-based COFs denoted TEBPY, TENPY, and TEPPY by a condensation reaction of hydrazine with 1,3,6,8-tetrakis(4-ethynylbenzaldehyde)-pyrene (TEBPY), 1,3,6,8-tetrakis(6-ethynylnicotinaldehyde)-pyrene (TENPY), and 1,3,6,8-tetrakis(2-ethynylpyrimidin-5-carb-aldehyde)pyrene (TEPPY), respectively (Fig. 17a).139
image file: d3ta02189k-f17.tif
Fig. 17 Synthesis and applications of pyrene-based COFs. (a) Synthesis scheme for pyrene-based azine-linked COFs. (b) Time course of photocatalytic hydrogen evolution from a 10 vol% aqueous TEOA suspension and Pt-modified A-TEXPY-COFs. Reproduced with permission from ref. 139 Copyright© 2018, Wiley-VCH.

The azine-linked COFs A-TEBPY, A-TENPY, and A-TEPPY showed water splitting activity with hydrogen generation rates of 98, 22, and 6 μmol h−1 g−1, respectively (Fig. 17b). The mechanism of photocatalytic water splitting was believed to follow a radical cation pathway, and the stability of the radical cation was found to increase with decreasing nitrogen content. As the nitrogen content increases in the COFs, the hydrogen generation rate was found to decrease, highlighting the effect of structural tuning at the molecular level on photocatalytic activity – a structure–property correlation.

In an attempt to create a pyrene-based COF donor–acceptor COF system, Jiang and coworkers developed a donor–acceptor-based sp2 carbon conjugated COF, named sp2c-COF, by solvothermal reaction between tetra(4-formylphenyl)pyrene (TFPPy) and 1,4-phenyldiacteonitrile (PDAN) and further functionalized it with 3-ethylrhodanine (ERDN) monomers as an end-capping group (Fig. 18a and b).140 These sp2c-COF samples showed excellent stability in water, acidic (HCl), and basic (NaOH) conditions. The COF showed good hydrogen evolution activity with a rate of 1360 μmol h−1 g−1 under visible light (≥420 nm) (Fig. 18c–e). Interestingly, COF functionalized with ERDN end groups showed enhanced hydrogen generation with a rate of 2120 μmol h−1 g−1 upon irradiation with a wavelength of 420 nm, which was found to be approximately 1.6 times higher than that of pristine COF. The enhanced hydrogen evolution activity of sp2c-COFERDN has been ascribed to its band structure, improved light-harvesting capability, and heterojunction formation. In addition, the electron-deficient nature of EDRN introduces a push–pull effect in the COF network, facilitating facile electron migration and thereby preventing charge recombination.


image file: d3ta02189k-f18.tif
Fig. 18 (a) Synthesis scheme of sp2-c-COF and sp2-c-COFEDRN. (b) Mechanism of hydrogen generation from water using a sp2 carbon-conjugated framework and Pt as a cocatalyst. (c) Apparent quantum yields (AQYs) of sp2c-COFERDN under irradiation with monochromatic light at 420 nm, 490 nm, 520 nm and 578 nm. (d) Hydrogen evolution was monitored over 5 h with sp2c-COF (blue circles), sp2c-COFERDN (red circles), sp2c-CMP (black circles) and imine-linked pyrene COF (triangles) as photocatalysts under irradiation with wavelengths ≥420 nm. (e) Long-term stability of sp2c-COFERDN over four repeated photocatalytic hydrogen evolution cycles under light irradiation (≥420 nm). Reproduced with permission from ref. 140 Copyright© 2019, Elsevier.

The design of thiazole-containing pyrene-based COFs (Tz-COF) exhibiting high crystallinity, surface areas, and good physicochemical stability has been reported by Cooper and coworkers.141 The COFs were prepared via C–H functionalization and subsequent oxidative annulation reaction of 2,6-diaminonaphthalene and 2,6-diaminoanthracene moieties (Fig. 19a). The one-pot synthesis of thiazole-based COFs was complete with thiazole ring (Tz) formation by a cascade reaction between aldehyde, amine, and sulfur involving the following steps: (i) Schiff base condensation between aldehyde and amine, (ii) electrophilic attack from sulfur and elimination Sn−1, (iii) cyclization, and (iv) oxidative aromatization of the ring. Photocatalytic investigations of crystalline Tz-COF-4 constituted by pyrene and thiazole building blocks displayed a hydrogen generation rate of 4296 μmol h−1 g−1 under visible light (λ > 420 nm) using ascorbic acid as a sacrificial electron donor, which was higher than that of the amorphous counterpart (Fig. 19b and c). The high hydrogen generation performance of Tz-COF-4 has been attributed to its large surface area and broad absorption of visible light.


image file: d3ta02189k-f19.tif
Fig. 19 (a) Scheme of the synthesis of thiazole-linked Tz-COFs. (b) Comparison of the absorption profile and (c) hydrogen evolution rate under visible light irradiation (λ > 420 nm) by Tz-COF-4, IL-COF-2 and amorphous Tz-COF-4. Reproduced with permission from ref. 141 Copyright© 2020, American Chemical Society.

In another precedent of the donor–acceptor strategy, Chen and coworkers developed pyrene-based isostructural Py-XTP-BT-COFs (X = H, F, Cl) by condensation of 4,4′,4′′,4′′′-(pyrene-1,3,6,8-tetrayl)tetrabenzaldehyde (Py-CHO) with terphenyl-based diamines, 4,7-diphenylbenzo[c][1,2,5]thiadiazole (HTP-BT-NH2) or 4,7-diphenyl-5,6-difluorobenzo[c][1,2,5]thiadiazole (FTP-BT-NH2) or 4,7-dibromo-5,6-dichlorobenzo[c][1,2,5]thiadiazole (ClTP-BT-NH2) (Fig. 20a).142 The water splitting studies showed higher hydrogen evolution activity of chlorinated COF (Py-ClTP-BT) with a rate of 177 μmol h−1 over fluorinated, Py-FTP-BT (57 μmol h−1) and nonhalogenated, Py-HTP-BT (21 μmol h−1) (Fig. 20b). The presence of electron-withdrawing halogen atoms on the benzothialozole linker facilitates effective charge separation by suppressing electron–hole recombination, lowering the activation barrier for hydrogen generation. This is also reflected in the average lifetimes of the photogenerated electrons of 3.8, 2.85, and 2.13 ns for Cl, F, and nonsubstituted COFs. This work represents another important demonstration of the effect of structural modulation on the photocatalytic property of COFs, providing a useful strategy for the rational design of COFs for enhanced hydrogen evolution from water.


image file: d3ta02189k-f20.tif
Fig. 20 (a) Scheme of synthesis of Py-XTP-BT COF (X = Cl, F, H). (b) Hydrogen evolution rates with and without Pt-cocatalysts. Reproduced with permission from ref. 142 Copyright© 2020, Wiley-VCH.

In a similar strategy, Chen and coworkers utilized functionalized benzothiadiazole (BT) and pyrene tetraamine building blocks to construct donor–acceptor photocatalytic NK-COFs through Schiff base condensation under solvothermal conditions (Fig. 21a).143 Notably, NK-COFs possess a high specific surface area and mesoporosity (∼35 Å) and exhibit good chemical stability in boiling water and acidic and basic solutions. The photocatalytic water splitting studies revealed higher hydrogen generation activity (120 μmol h−1) for NKCOF-108 composed of a fluorinated benzothiadiazole linker over the nonfluorinated counterpart NKCOF-110 (Fig. 21b). As discussed in the previous sections, the fluorination of benzothiadiazole enhances the electron-accepting nature of the linker, lowering the LUMO and HOMO levels and facilitating effective charge separation and migration from the electron-donor pyrene to the benzothiadiazole moiety, leading to enhanced hydrogen evolution performance.


image file: d3ta02189k-f21.tif
Fig. 21 (a) Scheme of synthesis of NK-COFs and (b) hydrogen evolution performance of various NK-COFs. Reproduced with permission from ref. 143 Copyright© 2021, American Chemical Society.

In a similar strategy, Wen and coworkers recently developed PyTz-COF by imine condensation between electron donor 4,4′,4′′,4′′′-(pyrene-1,3,6,8-tetrayl)tetraaniline (PyTA) and electron acceptor 4,4′-(thiazolo[5,4-d]thiazole-2,5-diyl) dibenzaldehyde (TzDA) moieties (Fig. 22a).144 To compare the photophysical properties of PyTz-COF with those of a non-donor-acceptor system, PyBp-COF was prepared, wherein the Tz linker was replaced by a biphenyl linker. The photoluminescence emission intensity of PyTz-COF was lower than that of PyBp-COF, suggesting effective charge separation in the former compared to that of the latter COF (Fig. 22b). The photocatalytic study of PyTz-COF for water splitting revealed an excellent hydrogen generation rate of 2072.4 μmol g−1 h−1 under simulated sunlight using ascorbic acid as a sacrificial electron donor (Fig. 22c). The enhanced HER of PyTz-COF has been ascribed to good visible light absorption and effective separation and transportation of charge carriers aided by the push–pull effect of the donor–acceptor COF.


image file: d3ta02189k-f22.tif
Fig. 22 (a) Preparation of donor–acceptor PyTz-COF. (b) Photoluminescence spectra of PyTz-COF and PyBp-COF excited at 320 nm. (c) Time-dependent photocatalytic hydrogen generation activity of PyTz-COF under AM 1.5 light irradiation. Reproduced with permission from ref. 144 Copyright© 2020 Wiley-VCH.

The controlled deposition of platinum (Pt) as a cocatalyst in the COF matrix is of paramount importance for photocatalytic hydrogen evolution, as the uncontrolled growth of large and non-uniform Pt nanoparticles limits the number of exposed active surfaces, even at higher loadings. To address this issue, Deng and co-workers synthesised a PY-DHBD-COF via a Schiff base condensation reaction involving 1,3,6,8-tetra(4-formylphenyl)pyrene (PY-CHO) and 1,4-dihydroxybenzidine (DHBD) (Fig. 23a).145 The resulting COF structure contained adjacent hydroxyl groups and imine bonds in its constituent units. The diffuse reflectance spectroscopy analyses showed that the COF material exhibited an absorption edge at ∼510 nm and the corresponding band gap was calculated to be 2.28 eV (Fig. 23b). The photocatalytic performance of the COF material for hydrogen generation under visible light irradiation was investigated using ascorbic acid as the sacrificial electron donor and Pt (H2PtCl6) as the cocatalyst. Remarkably, the highest observed photocatalytic hydrogen evolution rate reached 71[thin space (1/6-em)]160 μmol h−1 g−1 over 60 h when the Pt loading amount was 3.0 wt% with an AQY of 8.4% (Fig. 23c). During the photocatalytic hydrogen evolution, it was observed that the Pt cocatalyst exhibited selective adsorption near the hydroxyl groups and imine-N sites within the COF framework. Subsequent photoreduction processes resulted in the in situ photodeposition of highly dispersed Pt-clusters on the 2D layered surface of the COF. This study represents a significant advance in the field of photocatalysis by providing a new direction for precise control of co-catalyst deposition at the atomic level – contributing to the development of advanced strategies for improving the efficiency and effectiveness of photocatalytic processes.


image file: d3ta02189k-f23.tif
Fig. 23 Synthesis and photocatalytic activity of pyrene-based COF. (a) Schematic representation of the synthesis of PY-DHBD-COF. (b) Absorption spectra of PY-DHBD-COF [inset: Tauc plot]. (c) Recyclability of PY-DHBD-COF for hydrogen generation using 3 wt% Pt as cocatalyst. Reproduced with permission from ref. 145 Copyright© 2022, Springer Nature.

4.3 Porphyrin-based COFs

Porphyrins offer diverse applications in many fields, and the incorporation of porphyrins as building blocks in the construction of COF can significantly boost the chemical and photophysical properties of COF materials.146–148 Furthermore, the redox property of porphyrins can also play an essential role in hydrogen evolution.149 Porphyrin molecules possess a good visible light-harvesting ability, and the HOMO–LUMO gap can be engineered with suitable functionalization to design COFs with optimal properties for efficient water splitting for hydrogen generation. Hence, the design and application of porphyrin-based COFs as visible light-responsive catalysts for hydrogen evolution from water has gained significant interest from researchers worldwide.

Exfoliation has become a popular method for preparing layered materials since the discovery of exfoliated graphene. This method has enabled many promising electronic and optoelectronic applications of two-dimensional materials. To extend this advantage, the exfoliation of COFs has been attempted. However, due to strong hydrogen bonding and π–π stacking interactions among the layers, the exfoliation of COFs has limitations. To counter this problem, Osakada and coworkers proposed an exfoliation method for porphyrin COFs through the incorporation of metal ions (Mg and Cu) and an axial ligand (pyridine) to break the π–π stacking of the porphyrin COF via steric hindrance.150 For this study, a chemically stable porphyrin-based DhaTph COF was synthesized by reacting 5,10,15,20-tetrakis(4-aminophenyl)-21H,23H-porphyrin and 2,5-dihydroxyterephthalaldehyde in the presence of acetic acid. Using PXRD data, it was confirmed that the distance between two adjacent layers (001) was 4 Å, which approximately matches the simulated structure. Using simultaneous axial coordination of pyridines and metal ions, the synthesis of two-dimensional porphyrin nanodisks (e-CON) with an average thickness of 1 nm was achieved and further verified by scanning electronic microscopy. Considering the presence of extended π-conjugation and few-layered COF nanodisks, these materials were applied for photocatalytic hydrogen generation in the presence of Pt nanoparticles/reduced-graphene oxide. Enhanced hydrogen evolution was observed for e-CON (Cu,epy)/Pt/RGO compared with DhaTph/Pt/RGO, probably due to the increase in surface area between e-CON and Pt/RGO. This result demonstrated the role of COF exfoliation in enhancing the HER performance.

In a pioneering work, Wang and coworkers demonstrated the rational design of a series of porphyrin-based two-dimensional MPor-DETH-COFs (M = H2, Co, Ni, Zn) and investigated their photocatalytic hydrogen evolution activity.151 The COFs were synthesized by the Schiff base condensation of tetrakis(4-formylphenyl)pophyrin (p-MPor-CHO) and 2,5-diethoxy-terephthalohydrazide (DETH) under solvothermal conditions (Fig. 24a). The COFs showed good crystallinity and high specific surface areas in the range of 826–1020 m2 g−1 with pore channels of approximately 2.4 nm. UV-Vis absorption measurements of the COFs showed broad absorption in the visible region with narrow band gaps in the range of 1.77–1.88 eV and the longest excited-state lifetimes, indicating good charge-separation. In addition, the COFs showed good water-splitting activity under visible light in the presence of a Pt cocatalyst and triethanolamine as a sacrificial electron donor. The hydrogen generation rate of ZnPor-COF was found to be highest (413 μmol g−1 h−1), followed by NiPor- and H2Por-COF, with HERs of 211 and 80 μmol g−1 h−1, respectively. Notably, CoPor-COF showed the lowest hydrogen evolution rate of 25 μmol g−1 h−1 (Fig. 24b), which has been ascribed to the Co2+ (3d7) configuration with predominant LMCT that suppresses hole migration along the porphyrin channels. However, in the case of Ni2+ (3d8), the LMCT process is partially hampered, and as a result, hole transport is partly favored. On the other hand, for ZnPor-COF with Zn2+ (3d10), the LMCT process is forbidden, and hence, the photogenerated holes are free to migrate across the porphyrin rings and the electrons via the Zn⋯Zn chain, leading to the suppression of electron–hole recombination and resulting in higher hydrogen generation activity. Furthermore, the COFs showed good photostability and recyclability for multiple cycles of reuse, showcasing the utility of porphyrin-based COFs for photocatalytic water splitting for hydrogen generation.


image file: d3ta02189k-f24.tif
Fig. 24 (a) Schematic representation of the synthesis of BDF-TAPT-COFs; (b) hydrogen evolution rates for MPor-DETH-COFs (M = H2, Co, Ni, and Zn). Reproduced with permission from ref. 151 Copyright© 2021, Springer Nature.

4.4 Triphloroglucinol-containing COFs

After the first discovery of a COF by applying 1,3,5-triformylphloroglucinol (TP) as a linker,152 a large number of COFs have been prepared using the same linker and applied for various applications. The Schiff-base reaction of 1,3,5-triformylphloroglucinol with di-, tri- or tetra-amine linkers initiates through imine formation, followed by subsequent tautomerization to yield chemically stable β-ketoenamine-linked COFs. Because of their high chemical stability and conjugated structure, COFs constructed by using the covalent linking of 1,3,5-triformylphloroglucinol with other conjugated linkers have shown promising activity for water splitting. Note that the role of 1,3,5-triformylphloroglucinol in these COFs is to induce chemical stability and maintain conjugation, whereas the chemical functionalities and moieties (e.g., triazine, diacetylene, sulphone, etc.) induced through amine linkers initiate the water splitting reaction. The following section discusses the photocatalytic hydrogen generation activity of 1,3,5-triformylphloroglucinol-containing COFs.

In a major demonstration of the effect of conjugation on the water-splitting activity, we reported the effect of diacetylene moieties on the photocatalytic hydrogen generation activity of COFs.96 To analyze the role of diacetylene moieties in water splitting, β-ketoenamine-linked TP-EDDA and TP-BDDA COFs containing acetylene and diacetylene moieties, respectively, were prepared by solvothermal reaction of 1,3,5-triformylphloroglucinol with 4,4′-(ethyne-1,2-diyl)dianiline (EDDA) and/or 4,4′-(buta-1,3-diyne-1,4-diyl)dianiline (BDDA) linkers (Fig. 25a). The COFs exhibited good crystallinity, chemical stability and moderate surface areas (Fig. 25b). To further study the effect of conjugation on the optical and catalytic properties, an isoreticular TP-DTP COF was prepared by using 1,3,5-triformylphloroglucinol and 4,4′′-diamino-p-terphenyl (DTP) linkers. The UV-Vis spectra of the COFs showed an absorption edge at 500 nm for TP-DTP COF, whereas the absorption edges for TP-EDDA and TP-BDDA COF were redshifted to 520 and 525 nm, respectively (Fig. 25c). Interestingly, TP-BDDA COF showed better hydrogen evolution performance with a rate of 324 μmol h−1 g−1 over 60 hours, which was found to be approximately ten times higher than that of TP-EDDA COF (HER: 30 μmol h−1 g−1) and TP-DTP COF (HER: 20 μmol h−1 g−1). The enhanced hydrogen generation activity of TP-BDDA has been attributed to the presence of diacetylene moieties with extended π-conjugation. This resulted in facile separation and migration of photoexcited electrons to the catalyst surface for effective hydrogen generation. Using a series of COFs, we have demonstrated photocatalytic hydrogen generation using a COF-based photocatalyst without the presence of any heteronuclear functionalities (i.e., triazine, heptazine, etc.) inside the structure, emphasizing the importance of the diacetylene diad in photocatalysis.


image file: d3ta02189k-f25.tif
Fig. 25 Analyzing the role of diacetylene moieties in water splitting. (a) Scheme of the synthesis of TP-EDDA and TP-BDDA COFs. (b) Nitrogen sorption analyses of TP-EDDA and TP-BDDA COFs. (c) UV-Vis diffuse reflectance spectra of TP-BDDA, TP-EDDA, and TP-DTP COFs. (d) Photocatalytic hydrogen evolution performance of TP-EDDA, TP-BDDA and TP-DTP COFs. Reproduced with permission from ref. 96 Copyright© 2017, American Chemical Society.

In a significant development, Cooper and coworkers reported the design of sulfone-based S-COF (sulfone) and FS-COF (fused sulfone) and demonstrated the role of crystallinity, porosity and COF wettability on photocatalytic hydrogen generation activity. The synthesis of S-COF and FS-COF was achieved by a Schiff base condensation reaction between 1,3,5-triformylphloroglucinol and two planar thiophene-sulfone derivatives, 3,7-diaminodibenzo[b,d]thiophene sulfone (SA) and 3,9-diamino-benzo[1,2-b:4,5-b′]bis[1]benzothiophene (FSA), respectively (Fig. 26a).80 Owing to the irreversible keto–enol tautomerization of the products, the COFs formed were found to exhibit good chemical stability. Additionally, the effective π–π stacking between fused and planar FSA linkers facilitated the formation of highly crystalline FS-COF over S-COF. The photocatalytic water splitting studies showed impressive hydrogen evolution rates of 4.44 and 16.3 mmol h−1 g−1 by S-COF and FS-COF, respectively, under visible light (λ > 420 nm) irradiation with the Pt cocatalyst and ascorbic acid as a sacrificial electron donor (Fig. 26b). The photocatalytic activity was found to be consistent over 50 hours, validating the long-term durability of the COFs (Fig. 26c). It is noteworthy that the presence of sulphone moieties provided potential hydrophilic sites for enhancing the interaction/uptake of water by the COFs, thereby resulting in high water splitting activity.


image file: d3ta02189k-f26.tif
Fig. 26 Synthesis and applications of sulphone-based COFs. (a) Scheme of the synthesis of FS-COF. Inset: structures of 3,7-diaminodibenzo[b,d]thiophene 5,5-dioxide and [1,1′:4′,1′′-terphenyl]-4,4′′-diamine linkers. (b) Photocatalytic hydrogen evolution performance using FS-COF, S-COF, TP-COF and FS-P polymer. (c) Photocatalytic hydrogen generation under visible light irradiation using FS-COF over 50 h. Reproduced with permission from ref. 80 Copyright© 2018, Springer Nature.

In subsequent efforts, Lotsch and coworkers prepared a TpDTz COF utilizing 1,3,5-triformylphloroglucinol (TP) and 4,4′-(thiazolo[5,4-d]thiazole-2,5-diyl)dianiline (DTz) precursors and studied light-driven hydrogen evolution from water using a non-noble metal-based nickel-thiolate (NiME) cluster as a cocatalyst.153 Herein, the thiazolo[5,4-d]thiazole (TzTz) linkage provided high oxidative stability, excellent light absorbability, and high electron and hole mobility, rendering efficient photocatalytic performance to the COFs. Consequently, TpDTz COF exhibited good HER performance with a sustained hydrogen generation rate of 941 μmol g−1 h−1. Notably, the presence of the NiME cocatalyst was beneficial in increasing the COF excited-state lifetime during photocatalysis, resulting in better hydrogen evolution activity with long-term stability.

The modulation of active centers in the COF skeletons is of great importance for tuning the light absorption as well as photocatalytic water splitting performance, which, however, encounters severe challenges. In this regard, Chen and coworkers synthesized β-ketoenamine-linked Tp-PDA and Tp-DBN COFs through a Schiff base condensation reaction between 1,3,5-triformylphloroglucinol and phenylenediamine (PDA) or 2,5-diaminobenzonitrile (DBN) linkers, respectively (Fig. 27).98 Tp-PDA and Tp-DBN COFs exhibited hydrogen evolution with consistent rates of 0.6 mmol g−1 h−1 and 1.8 mmol g−1 h−1 H2 for up to 25 hours. The electron-withdrawing cyano (–CN) groups in Tp-DBN COF effectively participate in electron redistribution in the COF framework and develop efficient charge separation for photogenerated electron–hole pairs, further increasing the electron transfer rate during photocatalysis and enhancing the hydrogen evolution performance.


image file: d3ta02189k-f27.tif
Fig. 27 (a) Schematic representation of the synthesis method for Tp-PDA and Tp-DBN COFs. Hydrogen evolution activity (b) and incident light (wavelength)-dependent AQYs (c) for hydrogen generation using Tp-DBN COF carried out with 3 wt% Pt in sodium ascorbate solution. Reproduced with permission from ref. 98 Copyright© 2021, Wiley-VCH.

To establish the structure–property correlation, Le and coworkers synthesized several β-ketoenamine-based COFs, named TpPa-X [X = H, Cl2, SO3H and (CH3)2], by utilizing 1,3,5-triformylphloroglucinol and p-phenylenediamine (Pa-H), 2,5-dichloro-1,4-phenylenediamine (Pa-Cl2), 2,5-diaminobenzenesulphonic acid (Pa-SO3H) or 2,5-dimethyl-1,4-phenylenediamine (Pa-(CH3)2) (Fig. 28a).81 The photocatalytic investigations of the COFs revealed HER activity with rates of 11.13, 11.73, 4.44, and 3.62 μmol m−2 h−1 for TpPa-H, TpPa-Cl2, TpPa-SO3H, and TpPa-(CH3)2, respectively (Fig. 28b). The higher hydrogen evolution performance of TpPa-Cl2 has been correlated with the combined effects of optimal band structure, large specific surface area, better separation, and migration of charge carriers.


image file: d3ta02189k-f28.tif
Fig. 28 (a) Scheme of the synthesis of Tp-COFs using the solvothermal method. (b) The hydrogen evolution rate of different TpPa-H, TpPa-Cl2, TpPa-SO3H, and TpPa-(CH3)2, TpBD, and Tp-DTP COFs. Reproduced with permission from ref. 81 Copyright© 2021, Elsevier.

To modulate the photophysical properties of COFs via a bottom-up multicomponent design establishing the structure-to-activity relationships of multivariate COFs in terms of their photocatalytic hydrogen evolution, Guo and coworkers prepared a series of β-ketoenamine-based COFs by combining different molar ratios (x = 0, 0.05, 0.1, 0.25, 0.5, and 1) of 4,4′-(benzo-2,1,3-thiadiazole-4,7-diyl)dianiline (BT-NH2) and 4,4′-diamino-p-terphenyl (TP) linkers with 1,3,5-trifluoroformylglucinol.154 The photocatalytic studies of the COFs carried out under visible light (>420 nm) irradiation revealed the best water splitting performance by Tp(BT0.05TP0.95)-COF with a hydrogen evolution rate of 9839 μmol g−1 h−1. Herein, it was disclosed that the introduction of benzothiadiazole-derived units (chromophores) enhances the light-harvesting ability of COF, and due to its electron-withdrawing nature, effective separation and migration of photogenerated charge carriers was achieved, leading to improved photocatalytic hydrogen generation activity.

Recently, chirality has been shown to play a key role in determining the water splitting performance of materials through spin control. In further efforts, Guo and co-workers have investigated the potential of chiral β-ketoenamine-linked COF as artificial enzyme-like catalysts for visible photocatalytic hydrogen evolution. This work also presents a promising strategy for replacing platinum (Pt) as a cocatalyst in photocatalytic systems, thereby expanding the possibilities for developing efficient and sustainable artificial enzyme systems based on COF. In this contribution, achiral β-ketoenamine-linked COFs were synthesised by a solvothermal method involving aldimine condensation of 1,3,5-triformylphloroglucinol and diamine-substituted linkers (Fig. 29a).155 Subsequently, copper ion complexation was achieved by introducing Cu(OAc)2 into the aqueous dispersion of the COF (Fig. 29b). Photocatalytic hydrogen evolution in water was performed using the achiral TpPa-Cu(II)-COF photocatalyst under visible irradiation (λ > 420 nm) and L-cysteine as a sacrificial electron donor. Unfortunately, the TpPa-Cu(II)-COF with a copper loading of 10.76 wt% exhibited an inferior hydrogen evolution rate of 3640 μmol h−1 g−1 after 6 h of irradiation (Fig. 29c).


image file: d3ta02189k-f29.tif
Fig. 29 Investigation of the role of COF chirality in determining water splitting performance. (a) Synthesis scheme of achiral TpPa-Cu(II)-COF. (b) Structural tautomerism and Cu-complexation with a subunit. (c) Hydrogen evolution performances of achiral TpPa-Cu(II)-COF. (d) Comparison of hydrogen evolution performances for chiral COF using L/D-cystine as a sacrificial electron donor. (e) Comparison of turnover frequency between chiral and achiral COFs. Reproduced with permission from ref. 155 Copyright© 2022, Springer Nature.

Further, to explore the role of chirality in COFs for hydrogen evolution, the chiral TpPa-COF was synthesised using a chiral regulator to yield different enantiomeric forms: TpPa(Δ)-COF and TpPa(Λ)-COF, which were characterised using polarimetry and electronic circular dichroism. Based on these results, photocatalytic tests were carried out on the chiral TpPa-Cu(II)-COF using L- and D-cysteine as sacrificial electron donors. Remarkably, after 6 h of visible irradiation, the enantiomeric mixtures of TpPa(Δ)-Cu(II)-COF/L-cysteine and TpPa(Λ)-Cu(II)-COF/D-cysteine exhibited impressive hydrogen evolution rates of 14[thin space (1/6-em)]720 and 12[thin space (1/6-em)]800 μmol h−1 g−1, respectively (Fig. 29d), which are among the highest reported values for non-precious metal–organic photocatalysts. In comparison, the HER of TpPa(Δ)-Cu(II)-COF/L-cysteine was about four times higher than that of achiral TpPa-Cu(II)-COF/L-cysteine (3640 μmol h−1 g−1). In particular, the TpPa(Δ)-Cu(II)-COF/D-cysteine system achieved an HER of 10[thin space (1/6-em)]480 μmol h−1 g−1, whereas the TpPa(Λ)-Cu(II)-COF/L-cysteine system achieved an HER of 9480 μmol h−1 g−1 (Fig. 29e). The remarkable superiority of the chiral TpPa-COF can be attributed to its unique ability to allow parallel stacking of adjacent layers, with each atom in the upper layer aligned exactly with that in the lower layer. In contrast, the achiral TpPa-COF tends to stack its layers in an antiparallel fashion, as evidenced by the optimised layer structures. In the process of coupling the two adsorbed hydrogen (H) atoms to form H2, the proximity between the H atoms plays a crucial role in determining the kinetic favourability. Consequently, the chiral TpPa-COF, with its parallel layer stacking facilitating closer proximity of the H atoms, exhibits enhanced kinetic favourability in the H2 formation process, highlighting the significant influence of chirality on the photocatalytic performance.

4.5 Development of hybrid COFs for enhanced hydrogen generation

Hybrid materials are denoted as the most emergent materials at the edge of technological advancements, in which material properties are achieved through a synergetic combination of more than one component on the molecular scale. This makes hybrid materials interesting for several applications. It has been observed that the hydrogen generation activity of COFs can be enhanced significantly by hybridizing with a synergetic combination of different conductive or photoactive materials. In this context, the design of heterostructures composed of COFs and other materials, such as semiconductor metal oxides/sulfides, carbon-based materials, porous metal–organic frameworks (MOFs), etc., has been reported. In addition, the stable two-dimensional network of the COFs acts as a matrix to support the growth of nanoparticles. The heterojunctions formed favor the facile separation of the charge carriers by suppressing their recombination, resulting in higher photocatalytic efficiency. This section discusses various examples of COF-based hybrids reported for photocatalytic water splitting for hydrogen generation.
4.5.1 COF-metal oxide/sulfide-based hybrids. Cadmium sulfide (CdS) is a widely studied visible light active photocatalyst owing to its band gap (2.4 eV) and optical properties.156 However, CdS is normally coupled with other materials to overcome the limitations of photocorrosion and to improve its photocatalytic properties. In this direction, Banerjee and coworkers developed a CdS-COF hybrid by utilizing a stable TpPa-2 COF as a matrix for CdS nanoparticle growth and investigated the hybrids for hydrogen generation activity.107 TpPa-2 was obtained by a solvothermal reaction of 1,3,5-triformylphloroglucinol (Tp) and 3,6-dimethyl-1,4-diaminophenyl (Pa) as building blocks. The effect of varying the COF content on the photocatalytic efficiency of the hybrid was studied. Interestingly, the introduction of COF was found to increase the photocatalytic activity of the hybrid. Of the various compositions synthesized, the CdS-COF (90[thin space (1/6-em)]:[thin space (1/6-em)]10) hybrid with 10 wt% COF showed the best hydrogen evolution activity with a rate of 3678 μmol h−1 g−1, which is multiple times higher than that of the pure CdS catalyst (124 μmol h−1 g−1). The enhanced hydrogen generation activity of the hybrid was attributed to improved photostability and suppressed recombination of CdS photogenerated holes and electrons.

To enhance the photocatalytic hydrogen evolution efficiency, numerous efforts have been devoted. This includes the deposition of noble metals (e.g., Pt) and the addition of sacrificial reagents, which have been proven very effective. However, Pt is a rare and expensive noble metal; therefore, there is an interest in replacing Pt with a low-cost cocatalyst. In this context, the incorporation of a noble metal-free cocatalyst, Ni(OH)2, in the TpPa-2 COF matrix was studied by Dong coworkers to boost the HER performance of the COF.157 The various compositions of the hybrid, Ni(OH)2-X%/TpPa-2 (X: molar fraction), were prepared by in situ synthesis of Ni(OH)2 in the presence of a known amount of COF. Among the various composites prepared, Ni(OH)2–2.5%/TpPa-2 with 2.5 mol% Ni(OH)2 showed the best hydrogen generation activity with a rate of 1895 μmol h−1 g−1, which was found to be nearly 26 times higher than that of pure TpPa-2. The enhanced HER activity of the hybrid has been ascribed to the synergetic interaction between COF and Ni(OH)2. Additionally, the metallic Ni formed by in situ reduction of Ni(OH)2 facilitates the effective separation of photogenerated electron–hole pairs. Notably, the hydrogen evolution activity of the hybrid was nearly equivalent to that of 0.3 wt% Pt-loaded TpPa-2, highlighting the potential utility of Ni(OH)2 as a promising noble metal-free cocatalyst.

In these attempts, Zhang and coworkers designed Z-scheme heterostructured hybrids by embedding hematite (α-Fe2O3) onto TpPa-2 COF for effective visible-light-driven photocatalytic water splitting (Fig. 30a).158 The α-Fe2O3/TpPa-2 COF hybrid was prepared by adding α-Fe2O3 with the reaction of TpPa-2 COF precursors 1,3,5-triformylphloroglucinol and 3,6-dimethyl-1,4-diaminophenyl under solvothermal conditions. The photocatalytic studies of the hybrid showed the best water splitting activity for the α-Fe2O3/TpPa-2 COF (3[thin space (1/6-em)]:[thin space (1/6-em)]7) hybrid with a hydrogen evolution rate of 3.77 mmol h−1 g−1 in the absence of a noble metal cocatalyst (Fig. 30b and c). Interestingly, the activity was approximately 53 times higher than that of pristine TpPa-2-COF and higher than that of various reported Z-scheme photocatalysts. The higher HER performance of the hybrid has been assigned to its strong light-harvesting ability and formation of a Z-scheme heterojunction, which promotes efficient charge-carrier separation and migration for the reduction of protons to hydrogen. Later, Zhang and coworkers further extended this strategy for covalent anchoring of TiO2 nanostructures into TpPa-1 COF by a Schiff base reaction between surface-functionalized –CHO groups on TiO2 and –NH2 groups in the TpPa-1 COF reactant.159 Thus, various hybrids with varying TiO2[thin space (1/6-em)]:[thin space (1/6-em)]TpPa-1-COF ratios were prepared, and their HER activity was tested under visible light (420 nm) using Pt as a cocatalyst and sodium ascorbate as a sacrificial electron donor. Among the hybrids, TiO2-TpPa-1 COF (1[thin space (1/6-em)]:[thin space (1/6-em)]3) showed the best hydrogen evolution rate of 11.19 mmol h−1 g−1. Mechanistic studies revealed that the covalent bonds in the hybrid serve as a bridge in promoting electron–hole separation, enhancing the HER performance.


image file: d3ta02189k-f30.tif
Fig. 30 One-pot synthesis of the α-Fe2O3/TpPa-2-COF hybrid for water splitting. (a) Scheme of the synthesis of α-Fe2O3/TpPa-2-COFs. (b) Wavelength dependence of AQE for α-Fe2O3/TpPa-2-COF (3[thin space (1/6-em)]:[thin space (1/6-em)]7). (c) Hydrogen evolution rates of hybrids with different compositions. Reproduced with permission from ref. 158 Copyright© 2020, Royal Society of Chemistry.

The same research group further studied the integration of MoS2 with TpPa-1 COF to prepare a MoS2/TpPa-1 COF composite by in situ growth of COF in an exfoliated MoS2 dispersion (Fig. 31a).160 Various compositions were prepared by varying the loading of MoS2 from 1 to 5 wt% and further tested for photocatalytic water splitting. The optimized composite with 3 wt% loading of MoS2 exhibited a hydrogen evolution rate of 55.85 μmol h−1 without the use of a noble metal cocatalyst under visible light (Fig. 31d and e). Notably, the HER performance of the hybrid was 32 times higher than that of pristine TpPa-1 COF (1.72 μmol h−1). Herein, the introduction of MoS2 has been found to enhance the charge separation for photogenerated electron–hole pairs, resulting in enhanced HER performance (Fig. 31b). The mechanism involves the transfer of photoinduced electrons from the COF to MoS2 and their utilization for water reduction for hydrogen generation (Fig. 31c).


image file: d3ta02189k-f31.tif
Fig. 31 The preparation and applications of MoS2/TpPa-1-COF hybrids. (a) Synthetic route for the preparation of the MoS2/TpPa-1-COF hybrid. (b) EPR spectra of TpPa-1-COF and MoS2-3%/TpPa-1-COF with light on and off. (c) Schematic illustration of photocatalytic hydrogen evolution under visible light irradiation using the MoS2/TpPa-1-COF composite. (d) Photocatalytic hydrogen generation rates for TpPa-1 and MoS2/TpPa-1 COF composites with variable MoS2 content. (e) Comparison of the photocatalytic activity of TpPa-1-COF, MoS2-3%/TpPa-1-COF, Pt-3%/TpPa-1-COF and physically mixed (3%) material. Reproduced with permission from ref. 160 Copyright© 2019, Royal Society of Chemistry.

Similarly, CTF-based hybrid systems have been explored for photocatalytic water splitting. Since CTFs are generally prepared at high temperatures and using ZnCl2, the preparation of hybrids using these materials has not been explored much. In these attempts, Luo and coworkers applied a strategy to enhance the photocatalytic efficiency of CTF-1 by appending black phosphorus (BP) as a bridge joint to generate a sandwich-type Pt-containing CTF system (CTF-BP-Pt).161 Interestingly, CTF-BP-Pt exhibited a higher hydrogen evolution rate of 614.6 μmol h−1 g−1 over the CTF/Pt (167.5 μmol h−1 g−1) sample with directly doped Pt and the physical mixture of CTF-BP + Pt (95.1 μmol h−1 g−1). The superior hydrogen generation activity of CTF-BP-Pt has been attributed to the important role of black phosphorus nanosheets in enhancing the stability of Pt nanoparticles by effectively suppressing their aggregation.

CdS-based photocatalysts such as nanosized CdS powder, solid solutions and quantum dots have been widely explored for photocatalytic hydrogen generation due to their strong visible light absorption, suitable band edge levels and electronic charge transfer properties. Wu and coworkers reported a facile in situ growth of CdS quantum dots (QDs) on CTF-1 via a facile photoreduction method to further improve their hydrogen evolution performance in visible light.162 The CdS@CTF-1 composite displayed higher HER performance with a rate of 27.8 μmol h−1, which was approximately four times higher than that of pure CdS. The enhanced HER activity of the hybrid has been attributed to the synergistic interaction between CTF-1 and CdS, facilitating enhanced separation and transport of photoinduced charge carriers. In similar attempts, Wu and coworkers modified CTF-1 with reduced graphene oxide (rGO), and CTF-1/rGO-x hybrids with different loadings of rGO (x = 1, 2, 5, and 10 wt%) were obtained by using a facile UV reduction-based strategy.163 Interestingly, the hydrogen evolution rate of the composites was found to vary with the graphene oxide content, and the best hydrogen evolution activity with a rate of 894 μmol g−1 h−1 was observed for the CTF-1/rGO-2 composite, which was approximately 4.3-fold higher than that of pristine CTF.

Nanostructured transition metal phosphides (Ni, Co, Mo, W) have attracted high scientific interest due to their unique physical and chemical properties. To take advantage of those properties, Fan and coworkers developed a Ni2P/CTF composite and studied the effect of varying Ni2P loading on the hydrogen evolution performance.164 Notably, the rate of hydrogen evolution in the hybrid was found to increase gradually with an increase in the Ni2P loading and then decrease at higher loadings. The hybrid with 2.8% Ni2P loading showed the best hydrogen evolution activity with a rate of 5844 μmol g−1 h−1 and AQE of 3.8%. The higher hydrogen evolution performance of the hybrid has been attributed to the facile transfer of photogenerated electrons from CTF to Ni2P, resulting in effective charge separation and utilization. Interestingly, the hydrogen evolution rate of 2.8% Ni2P/CTF was equivalent to that of 3.0% Pt-loaded CTF, highlighting the importance of Ni2P as a promising noble-metal-free cocatalyst for enhanced visible light-assisted hydrogen generation.

4.5.2 MOF–COF hybrids. Recently, the hybridization of MOFs and COFs has attracted significant interest owing to the unique advantages of long-range ordering, high surface area, and excellent visible light harvesting properties of both MOFs and COFs.165 In this regard, Lan and coworkers developed a new class of MOF/COF hybrid materials by covalently anchoring NH2-UiO-66 on the surface of TpPa-1-COF and tested their photocatalytic hydrogen evolution.166 Herein, TpPa-1 COF was used to construct the composite, considering its rigid structure formed via Schiff-base linkage, and NH2-UiO-66 MOF was chosen due to its photocatalytic properties. The hybrid was prepared by introducing NH2-UiO-66 during the synthesis of TpPa-1 COF (Fig. 32a). Notably, the NH2-UiO-66/TpPa-1 COF hybrids feature a crystalline porous framework structure with a high surface area. Owing to the matching band gaps between COF and MOF, the hybrids exhibit efficient visible-light-assisted photocatalytic hydrogen evolution activity with a rate of 23.41 mmol g−1 h−1, which was approximately 20 times higher than that of pristine TpPa-1 COF (Fig. 32b). The higher photocatalytic efficiency of the hybrid has been attributed to the formation of a COF/MOF heterojunction that facilitates the transfer of photoinduced electrons from the COF to the MOF, at which the reduction of water to hydrogen takes place (Fig. 32c). It is worth highlighting that the hydrogen evolution activity of these MOF/COF hybrids was found to be superior to that of many previously reported examples of MOFs and COFs (Table 3).
image file: d3ta02189k-f32.tif
Fig. 32 Preparation of MOF-COF composites for water splitting. (a) Scheme of the synthesis of the NH2-UiO-66/TpPa-1-COF hybrid. (b) Hydrogen evolution performance of the NH2-UiO-66/TpPa-1-COF hybrid with different compositions. (c) Proposed mechanism for H2 generation using the NH2-UiO-66/TpPa-1-COF hybrid. Reproduced with permission from ref. 166 Copyright© 2018 Wiley-VCH.

In an interesting strategy, Yan and coworkers utilized postsynthetic modification of metal nodes from MOFs to prepare MOF/COF hybrids for photocatalytic water splitting. To prepare the hybrid, a Zr-MOF (MOF-808) without –NH2-containing ligands was first prepared and further modified by functionalization with p-aminobenzoic acid onto Zr6 clusters to anchor –NH2 groups.167 Then, the MOF-808@TpPa-1-COF core–shell structure was prepared by an in situ condensation reaction via covalent linkage between the COF precursors (Fig. 33a). Different MOF@COF hybrids were prepared by varying the weight ratios of the components. The hybrid possessing high crystallinity and hierarchical pore structure was applied for photocatalytic water splitting, where a hydrogen evolution activity as high as 11.88 mmol h−1 g−1 was revealed for the MOF-808@TpPa-1 COF (6/4) hybrid. This activity was approximately 5.6 times higher than that of pure TpPa-1 COF and the physical mixture (Fig. 33b and c). The higher photocatalytic efficiency of the hybrid has been correlated to facile separation and transport of photogenerated charge carriers at the heterojunction.


image file: d3ta02189k-f33.tif
Fig. 33 (a) Synthesis scheme of the MOF-808@TpPa-1-COF hybrid. (b and c) Hydrogen evolution performance of the hybrids. Reproduced with permission from ref. 167 Copyright© 2021, Royal Society of Chemistry.
4.5.3 COF-carbon-based hybrids. The application of graphitic carbon nitride (g-C3N4) as a photocatalyst has gained significant interest owing to its visible-light harvesting property (band gap of 2.7 eV), nontoxicity, low cost, good stability and metal-free nature.168 However, the photocatalytic activity of bare g-C3N4 is inhibited by a high charge recombination rate, which can be overcome by functionalization with suitable materials to form heterojunction photocatalysts with improved photocatalytic performance. In this regard, the first g-C3N4-modified CN-COF hybrid was prepared by in situ solvothermal reactions of g-C3N4 with 1,3,5-triformylphluoroglucinol (TP) and 4,4′,4′′-(1,3,5-triazine-2,4,6-triyl)trianiline (TTA) moieties.110 Different ratios of COF[thin space (1/6-em)]:[thin space (1/6-em)]CN were prepared, and their HER activity was investigated under visible light irradiation (λ > 420 nm). The optimized catalyst with 0.3 wt% COF showed the best hydrogen evolution rate of 10.1 mmol h−1 g−1, which was approximately 6.7 and 44.7 times higher than g-C3N4 and COF, respectively. The imine linkages between g-C3N4 and the COF material provide higher absorption in the visible region and reduce charge recombination with efficient charge transfer. Thus, the improved HER performance of the hybrid has been ascribed to enhanced broader absorption of visible light and effective separation and migration of photoexcited electron–hole pairs.

To construct heterojunction photocatalysts with multidimensional features, Li and coworkers designed hybrid photocatalysts composed of β-ketoenamine-based COF and g-C3N4 with varying mass ratios (1[thin space (1/6-em)]:[thin space (1/6-em)]x; x = 2.5, 5, 10, 15, 20).169 The composites were prepared via in situ reaction of 2,4,6-triformylphloroglucinol and benzidine building blocks in the presence of a stripped g-C3N4 suspension. The photocatalytic study of the composites showed the best HER activity by COF-CN (1[thin space (1/6-em)]:[thin space (1/6-em)]10) with a rate of 384 μmol h−1, which was found to be much higher than that of pure TpBD COF (1.35 μmol h−1) and g-C3N4 (6.24 μmol h−1). It is worth noting that the apparent quantum efficiency (AQE) for hydrogen generation by COF-CN (1[thin space (1/6-em)]:[thin space (1/6-em)]10) was approximately 15% (500 nm), which is the highest value among previously reported COFs or g-C3N4-based systems. The superior HER performance of the COF-CN hybrid has been attributed to the formation of a type-II heterojunction facilitating facile separation of photogenerated electron–hole pairs, resulting in enhanced photocatalytic activity. Similarly, Jiang and coworkers developed a COF/g-C3N4 hybrid by utilizing a donor–acceptor type TBTA COF through in situ syntheses by the condensation of 2,4,6-triformylphloroglucinol and 4,4′-(benzo-1,2,5-thiadiazole-4,7-diyl)dianiline (BTDA) in the presence of g-C3N4.170 The obtained hybrids possess optimal band edge positions suitable for effective water splitting and hence exhibit better photocatalytic activity than the individual components. Notably, the hybrid with the TBTA COF[thin space (1/6-em)]:[thin space (1/6-em)]g-C3N4 (2.5[thin space (1/6-em)]:[thin space (1/6-em)]1) composition showed the best activity with a hydrogen evolution rate of 11.73 mmol g−1 h−1 in the absence of a noble metal cocatalyst. Interestingly, the rate of hydrogen generation was further enhanced to 26.04 mmol g−1 h−1 with the addition of the Pt cocatalyst, highlighting the importance of hybrid systems for photocatalysis.

The construction of covalently bonded COF heterostructures has proven to be an effective way to improve photocatalytic HER performance. Toward the development of such heterostructures, the preparation of hybrids and Z-schemes using TiO2, g-C3N4, metal hydroxides, etc., have been attempted. Additionally, many hybrid materials have explored reduced graphene oxide (rGO) with high electron conductivity and a large surface area as an electron acceptor and mediator. To take advantage of the two-dimensional structure, porosity and conductivity of reduced graphene oxide, Zhang and coworkers developed a covalently connected rGO-COF (TpPa-1) hybrid photocatalyst for enhanced hydrogen evolution (Fig. 34a).108 The hybrids were prepared by following a one-pot synthetic route by taking a known amount of GO during the preparation of TpPa-1 COF using DMF solvent. During the synthesis, simultaneous reduction and covalent functionalization of GO with COF take place. Visible-light-assisted water splitting studies using these composites revealed that the rGO-TpPa-1 COF hybrid with a rGO composition of 5% showed the best hydrogen evolution rate of 11.98 mmol g−1 h−1, which was approximately 4.85 and 2.50 times higher than those of the pure TpPa-1-COF and 5%rGO/TpPa-1-COF physical mixture, respectively (Fig. 34b and c). The mechanistic investigation demonstrated that the covalent grafting of rGO with TpPa-1 COF in the rGO-TpPa-1 hybrid can act as an electron collector and transporter, promoting the separation of photogenerated charges and transferring the active electrons (Fig. 34d). In addition, the sheet morphology of the rGO-TpPa-1 COF hybrid with a uniform distribution of TpPa-1 on the rGO layers facilitates the migration of photogenerated electrons, demonstrating the importance of COF-based hybrids with strong bonds for efficient photocatalytic reactions.


image file: d3ta02189k-f34.tif
Fig. 34 Synthesis of COF-reduced graphene oxide hybrids for photocatalysis. (a) Scheme of the synthesis of the rGO-TpPa-1 COF hybrid. (b) The photocatalytic hydrogen evolution performance over five hours of a series of rGO (x%)-TpPa-1-COF hybrids. (c) Recycling test for photocatalytic hydrogen evolution of rGO (5%)-TpPa-1-COF. (d) Transient photocurrent response measurements of TpPa-1-COF and rGO (5%)-TpPa-1-COF vs. SCE. Reproduced from ref. 108 with permission Copyright© 2020, Royal Society of Chemistry.

Two-dimensional COFs with periodic structures, extended π-conjugation and layered stacking have emerged as promising materials for photocatalytic hydrogen evolution. However, the layer-by-layer assembly in 2D COFs is unstable during photocatalytic water splitting, resulting in disordered stacking and decreased HER performance. To address this concern, in a major development in the design of COF-based hybrids, Guo and coworkers prepared polyethylene glycol (PEG)-incorporated COF hybrids by postsynthetic modification of benzothiadiazole-containing BT-COF (Fig. 35a).171 The filling of the open channels of BT-COF with PEG led to the retention of the ordered stacking during the visible light-assisted photocatalytic reaction (Fig. 35b and c). This strengthening of interlayer π-stacking resulted in enhanced performance of PEG-filled BT-COF for water splitting and retained its structural stability over photocatalytic cycling (Fig. 35d). Thus, the PEG@BT-COF composite showed efficient photocatalytic performance with a hydrogen generation rate of 11.14 mmol h−1 g−1, which is higher than that of pure BT-COF (7.70 mmol h−1 g−1) (Fig. 35e and f). Notably, without PEG stuffing, the crystallinity and stability of the COF were reduced during photocatalysis, resulting in lower hydrogen evolution performance (Fig. 35e). Herein, Pt nanoparticles grown on PEG chains promote efficient photocatalytic performance. Thus, incorporating PEG in the one-dimensional pore channels of the COF enhanced the hybrid's stability and photocatalytic performance (Fig. 35g). This work highlights the importance of the rational design of COF-based hybrid catalysts to achieve enhanced photocatalytic water splitting efficiency.


image file: d3ta02189k-f35.tif
Fig. 35 Synthesis of PEG-stabilized 2D COFs. (a) Synthesis of BT-COF through a Schiff-base reaction. (b) HR-TEM image of BT-COF showing one-dimensional channels. (c) Scheme of transformation of COF and PEG@COF during Pt-cocatalyst deposition. (d) PXRD patterns for the as-prepared BT-COF, the Pt-deposited BT-COF before and after 48 h of cycling, and the recycled 30% PEG@BT-COF after 48 h of photocatalysis. (e) Photocatalytic hydrogen evolution under visible irradiation for BT-COF, BT-COF (HOAc), 30% PEG@BT-COF, poly(TpBT), and TP-COF. (f) The apparent quantum efficiency of BT-COF and 30%PEG@BT-COF. (g) Photocatalytic hydrogen generation using BT-COF and 30%PEG@BT-COF. Reproduced with permission from ref. 171 Copyright© 2021, Springer Nature.

5. Structure–property relationship among COF structures and photocatalytic performance

The use of photocatalysts for water splitting to hydrogen generation is the cornerstone of photocatalysis. Many efforts have been devoted to enhancing the light absorption, charge separation, and lifetime of charge carriers to overcome their recombination process to render improved photocatalytic efficiency. There are many important parameters, such as crystallinity, porosity, band gap, band positions, visible light absorption, surface wettability, charge transport, acidity, basicity, donor–acceptor structures, planarity, surface modification, chemical stability, functionalization, modular nature, presence of reactive sites and functional groups, structural designability, two-dimensional structure, π–π stacking, etc., govern the photocatalytic water splitting performance of the COFs (Fig. 36).
image file: d3ta02189k-f36.tif
Fig. 36 Most important properties responsible for determining the photocatalytic water splitting performance of COFs (generated using Word Cloud Generator by Monkey Learn).

Among these properties, some of the important parameters that determine the water-splitting performance and apparent quantum efficiency of COFs are as follows:

❖ Crystallinity: crystallinity is defined as the degree of long-range structural ordering comprising a crystal lattice within a material. The photocatalytic properties are strongly governed by the crystallinity of photocatalysts.178 An increase in the crystallite sizes has a strong impact on enhancing the photocatalytic rate, exhibiting low trap density and long diffusion length and thus demonstrating high catalytic performance. Highly crystalline cocatalysts show good photocatalytic performance due to their favorable charge transport properties.

❖ Porosity: porosity is a measure of the void spaces in a material, and a catalyst facilitates the exposure of active sites for effective photocatalytic hydrogen generation performance. It is well known that hierarchically porous structures are desirable for photocatalysis, as these materials provide more active sites, easy mass transport and shorten the carrier diffusion length between the carrier-generated centers and the active sites. Additionally, high porosity offers easy access to photoactive groups, solvents, sacrificial electron donors, cocatalysts, etc. – crucial components to achieve efficient water splitting.

❖ Band gap: generally, photocatalytic processes are initiated when a semiconductor photocatalyst absorbs photons with energy equal to or higher than its optical band gap. The photocatalytic water splitting performance of a photocatalyst strongly depends on its band-gap energy, which, in principle, should be smaller than 3 eV to extend the light absorption into the visible region, efficiently utilizing solar energy for hydrogen generation. In addition, the minimum band gap of semiconductor photocatalysts for water splitting should be 1.23 eV. Overall, the ideal band gap of COFs should be approximately 2.0 eV to effectively utilize solar energy.

❖ Band positions: in addition to the band gap, the optimum valence band and conduction band positions relative to the water redox potential are basic requirements of photocatalytic materials, principally determining the feasibility of hydrogen and oxygen evolution from water splitting under light illumination. Mainly, hydrogen and oxygen evolution reactions are known to occur when the conduction band edge of the photocatalyst is more negative than the reduction potential of H+/H2 and the valence band edge is more positive than the oxidation potential of O2/H2O.

❖ Surface wettability: surface wettability is a measurement of surface energy that influences the degree of contact with the physiologic environment. It is considered one of the critical parameters affecting the photocatalytic performance of COFs, which dominates the adsorption and desorption processes of reactants and products. The better surface wettability of COFs with water helps to enhance the dispersion of materials (e.g., sulfonyl COFs) and interaction among the active sites (e.g., ether or hydroxyl groups) present in the COF matrix with water, increasing the photocatalytic water splitting performance for hydrogen generation many-fold.

❖ Donor–acceptor structures: donor–acceptor (D–A) materials are constructed using an alternating array of donor and acceptor moieties. Two-dimensional COFs with π–π stacking among the adjacent layers can promote exciton migration and charge transport through the aligned one-dimensional π-columnar arrays, exhibiting attractive promise in photocatalytic applications. However, the photocatalytic performances are generally obstructed by insufficient charge separation and fast charge carrier recombination. To overcome these challenges, the design of donor–acceptor COFs with an intrinsic heterostructure of spatially separated donor and acceptor columns is essential. These D–A systems support effective charge separation with high electron mobility by facile transfer of photoexcited electrons from the donor molecule to the acceptor. However, strong donor and acceptor moieties are essential to boost facile charge migration. Hence, rational construction of donor–acceptor structures offers an intriguing option to enhance the photocatalytic hydrogen generation efficiency of COFs.

❖ Presence of photoactive moieties: by integrating photoactive moieties, the pore structures and electronic properties of two-dimensional COFs can be tuned, which can be further employed to enhance the photocatalytic properties of the COFs. To achieve efficient light absorption, generation, and separation/migration of excitons to the photocatalyst surface, the integration of various functional groups, donor–acceptor units, and photoactive moieties has been attempted. In these regards, generally, the light absorption can be extended toward the visible or infrared region by incorporating photoactive moieties such as triazine, heptazine, pyrene, acetylene, diacetylene, sulfonyl, and benzothiadiazole.

❖ Functionalization: the modular synthetic routes used for COFs allow the incorporation of suitable cocatalysts such as metals, metal oxide/sulfide and metal phosphide to achieve enhanced HER performance. Furthermore, the construction of heterojunctions, such as type II or Z-scheme systems, is a promising approach to facilitate enhanced photocatalytic hydrogen generation, aided by the easy separation of photogenerated carriers at the interface.

6. Summary and perspective

In summary, the literature covered in this review highlighted the importance of COF-based photocatalysts for visible-light-driven photocatalytic hydrogen generation. The modular nature of these materials facilitates rational incorporation of desired functionalities by judicious choice of organic linkers to obtain COFs with high surface area, optimal absorption of visible light and charge transport properties suitable for enhanced water splitting activity (Table 1). Furthermore, band gap engineering of COFs is a useful strategy to tailor the photocatalytic properties of COFs. Hence, the nature of the organic linkers employed will guide the structure and function of the resulting COFs and their photocatalytic performance toward efficient water splitting for hydrogen generation (Table 2). Thus, understanding the structure–property correlation is essential in developing COFs with promising photocatalytic properties for hydrogen evolution from water. From the aforementioned discussion, it is clear that the poor hydrogen generation performance of COFs stems from the loss of photogenerated electron–hole pairs by recombination. In this regard, the application of cocatalysts was proven to be essential to suppress this recombination of charge carriers and to facilitate facile migration of photogenerated electrons to the cocatalyst for effective hydrogen generation. In this regard, COFs composed of extended π-conjugation and donor–acceptor type linkers are promising for facilitating facile transport of photogenerated electrons by a push–pull mechanism from the donor moiety to the acceptor molecule and then effective transfer to the cocatalyst for the reduction of protons for hydrogen evolution.

Another major bottleneck that hampers photocatalytic efficiency is the loss of photogenerated charge carriers due to the presence of surface defects and imperfections. To overcome these challenges, the COF materials employed should possess good crystallinity, yielding defect-free and fewer grain boundaries. In this regard, the synthetic method employed to prepare COFs plays a vital role in generating COFs with high crystallinity. As discussed in the previous section, the COFs synthesized using the solvothermal route are found to possess high crystallinity over those prepared by other synthetic methods.

In addition, optimal absorption of light in the broader region of the solar spectrum covering the UV-Vis to near-infrared region (NIR) is recommended to achieve enhanced photocatalytic efficiency. However, most of the COFs reported thus far are active mainly in the UV-Vis region; hence, the future design of COFs should consider incorporating suitable functionalities/chromophores capable of harvesting light in the near-infrared (NIR) region. The fabrication of heterojunction photocatalysts can achieve this by coupling COFs with other suitable materials to generate composites with improved light-harvesting properties in the broad region of sunlight (Table 3). Furthermore, the COFs employed for photocatalysis should exhibit high photostability and long-term durability for sustainable water splitting to hydrogen evolution. In this context, heterojunction materials constructed by covalent attachment of COFs with 2D carbon-based materials such as g-C3N4, rGO, and graphene are found to be promising candidate materials for accomplishing improved catalytic efficiency.44,172 The resulting 2D/2D heterostructures possess promising features of improved light absorption, facile charge separation and migration aided by heterojunction formation. In addition, covalent functionalization provides extra stability to the COF network, rendering high photostability and long-term durability for multiple photocatalysis cycles. Recently, hybrids composed of COFs and MOFs have also gained significant interest, as they incorporate unique features of crystallinity, porosity, and light harvesting essential for an effective photocatalytic system. Hence, the future design of COFs should consider heterogenization with suitable materials for improved photocatalytic performance for the HER.173–175 However, it is worth mentioning that the best hydrogen generation activity of COFs has been achieved thus far in the presence of noble metal platinum cocatalysts, which limits their large-scale applications. Thus, the development of COFs exhibiting high H2 generation activity with non-Pt metal-based cocatalysts is highly desired. In this direction, the application of non-noble metal-based oxides, sulfides and phosphides with superior HER performance offers potential significance and should be considered in the future design of COF-based photocatalysts for hydrogen generation.176

Despite considerable efforts in designing various COF-based photocatalysts for hydrogen generation, the highest hydrogen evolution rate achieved thus far is not the best compared to conventional semiconductor materials. Furthermore, for practical applications of hydrogen as a clean energy source, the solar-to-fuel conversion efficiency needs to be enhanced along with the long-term durability of the catalyst. In addition, for large-scale implementation of hydrogen generation, several factors, such as the cost of the COF material, suitable reactor design, optimal exposure of sunlight on the photocatalyst, and facile separation of the catalyst from the reaction mixture, need to be optimized. In this regard, recent literature reports have demonstrated the application of planar panels of photocatalytic materials by coating powder material on suitable supports for large-scale hydrogen evolution applications.177 To achieve large-scale hydrogen generation, the large-scale synthesis of COFs with high crystallinity and porous structure needs to be optimized. Unfortunately, the high cost of large-scale preparation of COF semiconducting materials limits photocatalytic hydrogen production on an industrial scale. In this respect, the use of building blocks derived from natural sources or low-cost chemicals (e.g. urea, melamine, etc.) can reduce the overall cost of COF photocatalysts. In addition, to further reduce the cost, a photocatalytic water splitting system operating without a cocatalyst and sacrificial electron donor would be ideal. Moreover, precise knowledge of the thermodynamic aspects of the photocatalytic hydrogen generation mechanism can provide deep insight into the problem that can be helpful for the development of highly efficient COFs. Overall, the COFs and COF-based hybrid materials show great promise as photocatalysts for efficient visible light-mediated hydrogen evolution, and it is anticipated that future research in this area will benefit the advancement of COFs with optimal properties for large-scale photocatalytic water splitting to achieve sustainable generation of green hydrogen and toward the realization of a hydrogen economy.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

K. P. acknowledges IIT Ropar for the institute's postdoctoral fellowship. C.M.N. acknowledges IIT Ropar and DST-SERB (grant no. CRG/2022/006762). P. P. acknowledges the financial support from S. N. Bose National Centre for Basic Sciences and the Department of Science & Technology (No. SRG/2022/000217). P. P. and D. D. D. thank the Spanish Ministry of Science and Innovation for the project TED2021-132847B-100. D. D. D. thanks NANOtec, INTech, Cabildo de Tenerife and ULL for laboratory facilities. P. P. and B. M. would also like to thank Prof. Tanusri Saha-Dasgupta (Director) and Prof. Rajib Kumar Mitra (Head of Department) for their support, and Prof. Goutam De for the scientific discussion.

References

  1. K. Calautit, D. S. N. M. Nasir and B. R. Hughes, Renewable Sustainable Energy Rev., 2021, 147, 111230 CrossRef.
  2. R. Cassia, M. Nocioni, N. C. Aragunde and L. Lamattina, Front. Plant Sci., 2018, 9, 473 CrossRef PubMed.
  3. O. Ellabban, H. A. Rub and F. Blaabjerg, Renewable Sustainable Energy Rev., 2014, 39, 748–764 CrossRef.
  4. M. Kokkonen, P. Talebi, J. Zhou, S. Asgari, S. A. Soomro, F. Elsehrawy, J. Halme, S. Ahmad, A. Hagfeldt and S. G. Hashmi, J. Mater. Chem. A, 2021, 9, 10527–10545 RSC.
  5. J. Y. Kim, J. W. Lee, H. S. Jung, H. Shin and N.-G. Park, Chem. Rev., 2020, 120, 7867–7918 CrossRef CAS PubMed.
  6. K. Rogdakis, N. Karakostas and E. Kymakis, Energy Environ. Sci., 2021, 14, 3352–3392 RSC.
  7. Y. Ma, Y. Ma, Q. Wang, S. Schweidler, M. Botros, T. Fu, H. Hahn, T. Brezesinski and B. Breitung, Energy Environ. Sci., 2021, 14, 2883–2905 RSC.
  8. E. B. Agyekum, C. Nutakor, A. M. Agwa and S. Kamel, Membranes, 2022, 12, 173 CrossRef CAS PubMed.
  9. M. Yue, H. Lambert, E. Pahon, R. Roche, S. Jemei and D. Hissal, Renewable Sustainable Energy Rev., 2021, 146, 111180 CrossRef.
  10. J. Luo, S. Zhang, M. Sun, L. Yang, S. Luo and J. C. Crittenden, ACS Nano, 2019, 13, 9811–9840 CrossRef CAS PubMed.
  11. C. Acar, I. Dincer and G. F. Naterer, Int. J. Energy Res., 2016, 40, 1449–1473 CrossRef CAS.
  12. C. Acer and I. Dincer, J. Cleaner Prod., 2019, 218, 835–849 CrossRef.
  13. J. Qi, W. Zhang and R. Cao, Adv. Energy Mater., 2018, 8, 1701620 CrossRef.
  14. W. Liu, Y. Cui, X. Du, Z. Zhang, Z. Chao and Y. Deng, Energy Environ. Sci., 2016, 9, 467–472 RSC.
  15. A. Alnouss, G. McKay and T. Al-Ansari, Appl. Energy, 2020, 266, 144885 CrossRef.
  16. M. Watanabe, H. Inomata and K. Arai, Biomass Bioenergy, 2021, 22, 405–410 CrossRef.
  17. J. V. Karaeva, Int. J. Hydrogen Energy, 2021, 46, 34089–34096 CrossRef CAS.
  18. A. Fujishima and K. Honda, Nature, 1972, 238, 37–38 CrossRef CAS PubMed.
  19. Y. P. Xie, G. S. Wang, E. L. Zhang and X. Zhang, Chin. J. Inorg. Chem., 2017, 33, 177–209 CAS.
  20. M. Humayun, F. Raziq, A. Khan and W. Luo, Green Chem. Lett. Rev., 2018, 11, 86–102 CrossRef CAS.
  21. N. Cheng, S. Stambula, D. Wang, M. N. Banis, J. Liu, A. Riese, B. Xiao, R. Li, T.-K. Sham, L.-M. Liu, G. A. Botton and X. Sun, Nat. Commun., 2016, 7, 13638 CrossRef CAS.
  22. H. Nishiyama, T. Yamada, M. Nakabayashi, Y. Maehara, M. Yamaguchi, Y. Kuromiya, Y. Nagatsuma, H. Tokudome, S. Akiyama, T. Watanabe, R. Narushima, S. Okunaka, N. Shibata, T. Takata, T. Hisatomi and K. Domen, Nature, 2021, 598, 304–307 CrossRef CAS PubMed.
  23. Q. Wang, M. Nakabayashi, T. Hisatomi, S. Sun, S. Akiyama, Z. Wang, Z. Pan, X. Xiao, T. Watanabe, T. Yamada, N. Shibata, T. Takata and K. Domen, Nat. Mater., 2019, 18, 827–832 CrossRef CAS.
  24. A. Ishikawa, T. Takata, T. Matsumura, J. N. Kondo, M. Hara, H. Kobayashi and K. Domen, J. Phys. Chem. B, 2004, 108, 2637–2642 CrossRef CAS.
  25. A. Ishikawa, T. Takata, J. N. Kondo, M. Hara, H. Kobayashi and K. Domen, J. Am. Chem. Soc., 2002, 124, 13547–13553 CrossRef CAS PubMed.
  26. G. Liao, Y. Gong, L. Zhang, H. Gao, G.-J. Yang and B. Fang, Energy Environ. Sci., 2019, 12, 2080–2147 RSC.
  27. V. W.-h. Lau, M. B. Mesch, V. Duppel, V. Blum, J. Senker and B. V. Lotsch, J. Am. Chem. Soc., 2015, 137, 1064–1072 CrossRef CAS PubMed.
  28. J. A. Nasir, Z. U. Rehman, S. N. A. Shah, A. Khan, I. S. Butler and C. R. A. Catlow, J. Mater. Chem. A, 2020, 8, 20752–20780 RSC.
  29. B. Debnath, S. Dhingra and C. M. Nagaraja, Sol. RRL, 2021, 5, 2100226 CrossRef CAS.
  30. S. Dhingra, M. Sharma, V. Krishnan and C. M. Nagaraja, J. Colloid Interface Sci., 2022, 615, 346–356 CrossRef CAS.
  31. M. Kaur and C. M. Nagaraja, ACS Sustainable Chem. Eng., 2017, 5, 4293–4303 CrossRef CAS.
  32. P. Wu, J. Wang, J. Zhao, L. Guo and F. E. Osterloh, Chem. Commun., 2014, 50, 15521–15524 RSC.
  33. X. Wang, K. Maeda, X. Chen, K. Takanabe, K. Domen, Y. Hou, X. Fu and M. Antonietti, J. Am. Chem. Soc., 2009, 131, 1680–1681 CrossRef CAS PubMed.
  34. L. Shen, Z. Xing, J. Zou, Z. Li, X. Wu, Y. Zhang, Q. Zhu, S. Yang and W. Zhou, Sci. Rep., 2017, 7, 41978 CrossRef CAS.
  35. T.-X. Wang, H.-P. Liang, D. A. Anito, X. Ding and B.-H. Han, J. Mater. Chem. A, 2020, 8, 7003–7034 RSC.
  36. Y. Deng, Y. Wang, Y. Chen and Z. Zhang, Chem.–Asian J., 2021, 16, 1851–1863 CrossRef CAS PubMed.
  37. H. Kim, N. Kim and J. Ryu, Inorg. Chem. Front., 2021, 8, 4107–4148 RSC.
  38. T. Zhang, G. Xing, W. Chen and L. Chen, Mater. Chem. Front., 2020, 4, 332–353 RSC.
  39. L. Tan and B. Tan, Chem. Soc. Rev., 2017, 46, 3322–3356 RSC.
  40. N. B. McKeown, S. Makhseed and P. M. Budd, Chem. Commun., 2002, 2780–2781 RSC.
  41. Y. Yuan and G. Zhu, ACS Cent. Sci., 2019, 5, 409–418 CrossRef CAS PubMed.
  42. M. Liu, L. Guo, S. Jin and B. Tan, J. Mater. Chem. A, 2019, 7, 5153–5172 RSC.
  43. A. P. Côté, A. I. Benin, N. W. Ockwig, M. O'Keeffe, A. J. Matzger and O. M. Yaghi, Science, 2005, 310, 1166–1170 CrossRef PubMed.
  44. C. S. Diercks and O. M. Yaghi, Science, 2017, 355, 923–931 CrossRef CAS PubMed.
  45. S. Yanagida, A. Kabumoto, K. Mizumoto, C. Pac and K. Yoshino, J. Chem. Soc., Chem. Commun., 1985, 474–475 RSC.
  46. D.-L. Jiang, C.-K. Choi, K. Honda, W.-S. Li, T. Yuzawa and T. Aida, J. Am. Chem. Soc., 2004, 126, 12084–12089 CrossRef CAS PubMed.
  47. M. G. Schwab, M. Hamburger, X. Feng, J. Shu, H. W. Spiess, X. Wang, M. Antoniettib and K. Müllen, Chem. Commun., 2010, 46, 8932–8934 RSC.
  48. S. Haldar, M. Wang, P. Bhauriyal, A. Hazra, A. H. Khan, V. Bon, M. A. Isaacs, A. De, L. Shupletsov, T. Boenke, J. Grothe, T. Heine, E. Brunner, X. Feng, R. Dong, A. Schneemann and S. Kaskel, J. Am. Chem. Soc., 2022, 20, 9101–9112 CrossRef PubMed.
  49. N. Haung, P. Wang and D. Jiang, Nat. Rev. Mater., 2016, 1, 16068 CrossRef.
  50. Z. Wang, L. Liu, Z. Li, N. Goyal, T. Du, J. He and G. K. Li, Energy Fuels, 2022, 6, 2927–2944 CrossRef.
  51. B. He, Q. Zhang, Z. Pan, L. Li, C. Li, Y. Ling, Z. Wang, M. Chen, Z. Wang, Y. Yao, Q. Li, L. Sun, J. Wang and L. Wei, Chem. Rev., 2022, 11, 10087–10125 CrossRef PubMed.
  52. X. Guan, Q. Fang, Y. Yan and S. Qiu, Acc. Chem. Res., 2022, 55, 1912–1927 CrossRef CAS PubMed.
  53. K. T. Tan, S. Ghosh, Z. Wang, F. Wen, D. Rodríguez-San-Miguel, J. Feng, N. Huang, W. Wang, F. Zamora, X. Feng, A. Thomas and D. Jiang, Nat. Rev. Methods Primers, 2023, 3, 1 CrossRef CAS.
  54. R. Liu, K. T. Tan, Y. Gong, Y. Chen, Z. Li, S. Xie, T. He, Z. Lu, H. Yanga and D. Jiang, Chem. Soc. Rev., 2021, 50, 120–242 RSC.
  55. T. Skorjanc, D. Shetty, F. Gándara, Li. Ali, J. Raya, G. Das, M. A. Olson and A. Trabolsi, Chem. Sci., 2020, 11, 845–850 RSC.
  56. T. Zhou, Y. Zhao, J. W. Choi and A. Coskun, Angew. Chem., 2019, 131, 16951–16955 CrossRef.
  57. B. Gui, G. Lin, H. Ding, C. Gao, A. Mal and C. Wang, Acc. Chem. Res., 2020, 53, 2225–2234 CrossRef CAS PubMed.
  58. V. Singh and H. R. Byon, Mater. Adv., 2021, 2, 3188–3212 RSC.
  59. K. Geng, T. He, R. Liu, S. Dalapati, K. T. Tan, Z. Li, S. Tao, Y. Gong, Q. Jiang and D. Jiang, Chem. Rev., 2020, 120, 8814–8933 CrossRef CAS PubMed.
  60. J. Sprachmann, T. Wachsmuth, M. Bhosale, D. Burmeister, G. J. Smales, Z. Kochovski, N. Grabicki, B. Esser and O. Dumele, ChemRxiv. Cambridge, Cambridge Open Engage, 2022,  DOI:10.26434/chemrxiv-2022-p0pjr.
  61. D. Jiang, Bull. Chem. Soc. Jpn., 2021, 94, 1215–1231 CrossRef CAS.
  62. I. Ahmed and S. H. Jhung, Coord. Chem. Rev., 2021, 441, 213989 CrossRef CAS.
  63. S. Ghosh, A. Nakada, M. A. Springer, T. Kawaguchi, K. Suzuki, H. Kaji, I. Baburin, A. Kuc, T. Heine, H. Suzuki, R. Abe and S. Seki, J. Am. Chem. Soc., 2020, 142, 9752–9762 CAS.
  64. T. Banerjee, K. Gottschling, G. Savasci, C. Ochsenfeld and B. V. Lotsch, ACS Energy Lett., 2018, 3, 400–409 CrossRef CAS PubMed.
  65. M. Z. Rahman, Md. G. Kibria and C. B. Mulllins, Chem. Soc. Rev., 2020, 49, 1887–1931 RSC.
  66. M. G. Walter, E. L. Warren, J. R. McKone, S. W. Boettcher, Q. Mi, E. A. Santori and N. S. Lewis, Chem. Rev., 2010, 110, 6443–6445 CrossRef PubMed.
  67. N. Romero, R. Bofill, L. Francàs, J. G. Antón and X. Sala, Catalysts, 2021, 11, 754 CrossRef CAS.
  68. Q. Wang and K. Domen, Chem. Rev., 2020, 120, 919–985 CrossRef CAS PubMed.
  69. J. Li, D. Zhao, J. Liu, A. Liu and D. Ma, Molecules, 2020, 25, 2425 CrossRef CAS PubMed.
  70. H. Shen, W. Zhou, F. He, Y. Gu, Y. Li and Y. Li, J. Mater. Chem. A, 2020, 8, 4850–4855 RSC.
  71. A. M. Elewa, M. H. Elsayed, A. F. M. EL-Mahdy, C.-L. Chang, L.-Y. Ting, W.-C. Lin, C.-Y. Lu and H.-H. Chou, Appl. Catal., B, 2021, 285, 119802 CrossRef CAS.
  72. M. Natali, R. Argazzi, C. Chiorboli, E. Iengo and F. Scandola, Chem.–Eur. J., 2013, 19, 9261–9271 CrossRef CAS PubMed.
  73. H. Ozawa, M.-A. Haga and K. Sakai, J. Am. Chem. Soc., 2006, 128, 4926–4927 CrossRef CAS PubMed.
  74. W. Zhang, Y. Wang, Z. Wang, Z. Zhong and R. Xu, Chem. Commun., 2010, 46, 7631–7633 RSC.
  75. N. Fajrina and M. Tahir, Int. J. Hydrogen Energy, 2019, 44, 540–577 CrossRef CAS.
  76. J. You, Y. Zhao, L. Wang and W. Bao, J. Cleaner Prod., 2021, 291, 125822 CrossRef CAS.
  77. L. Stegbauer, K. Schwinhghammer and B. V. Lotsch, Chem. Sci., 2014, 5, 2789–2793 RSC.
  78. V. S. Vyas, F. Haase, L. Stegbauer, G. Savasci, F. Podjaski, C. Ochsenfeld and B. V. Lotsch, Nat. Commun., 2015, 6, 8508 CrossRef CAS PubMed.
  79. T. Banerjee, F. Haase, G. Savasci, K. Gottschling, C. Ochsenfeld and B. V. Lotsch, J. Am. Chem. Soc., 2017, 139, 16228–16234 CrossRef CAS PubMed.
  80. X. Wang, L. Chen, S. Y. Chong, M. A. Little, Y. Wu, W.-H. Zhu, R. Clowes, Y. Yan, M. A. Zwijnenburg, R. S. Sprick and A. I. Cooper, Nat. Chem., 2018, 10, 1180–1189 CrossRef CAS PubMed.
  81. L. Yin, Y. Zhao, Y. Xing, H. Tan, Z. Lang, W. Ho, Y. Wang and Y. Li, Chem. Eng. J., 2021, 419, 129984 CrossRef CAS.
  82. G.-B. Wang, S. Li, C. –X. Yan, Q. –Q. Lin, F. –C. Zhu, Y. Geng and Y. –B. Dong, Chem. Commun., 2020, 56, 12612–12615 RSC.
  83. J. Yang, A. Acharjya, M.-Y. Ye, J. Rabeah, S. Li, Z. Kochovski, S. Youk, J. Roeser, J. Grgneberg, C. Penschke, M. Schwarze, T. Wang, Y. Lu, R. v. d. Krol, M. Oschatz, R. Schomacker, P. Saalfrank and A. Thomas, Angew. Chem., Int. Ed., 2021, 60, 19797–19803 CrossRef CAS PubMed.
  84. C. Li, J. Liu, H. Li, K. Wu, J. Wang and Q. Yang, Nat. Commun., 2022, 13, 2357 CrossRef CAS PubMed.
  85. M. Ai, J.-W. Zhang, Y.-W. Wu, L. Pan, C. Shi and J.-J. Zou, Chem.–Eur. J., 2021, 27, 13669–13698 CrossRef PubMed.
  86. X. Li, C. Yang, B. Sun, S. Cai, Z. Chen, Y. Lv, J. Zhang and Y. Liu, J. Mater. Chem. A, 2020, 8, 16045–16060 RSC.
  87. Y. Li, W. Chen, G. Xing, D. Jiang and L. Chen, Chem. Soc. Rev., 2020, 49, 2852–2868 RSC.
  88. H. Wei, S. Chai, N. Hu, Z. Yang, L. Wei and L. Wang, Chem. Commun., 2015, 51, 12178–12181 RSC.
  89. X. Guan, Y. Ma, H. Li, Y. Yusran, M. Xue, Q. Fang, Y. Yan, V. Valtchev and S. Qiu, J. Am. Chem. Soc., 2018, 140, 4494–4498 CrossRef CAS PubMed.
  90. A. d. l. P. Ruigómez, D. R.-S. Miguel, K. C. Stylianou, M. Cavallini, D. Gentili, F. Liscio, S. Milita, O. M. Roscioni, M. L. R. González, C. Carbonell, Da. Maspoch, R. Mas-Ballesté, J. L. Segura and F. Zamora, Chem.–Eur. J., 2015, 21, 10666–10670 CrossRef PubMed.
  91. M. Matsumoto, R. R. Dasari, W. Ji, C. H. Feriante, T. C. Parker, S. R. Marder and W. R. Dichtel, J. Am. Chem. Soc., 2017, 139, 4999–5002 CrossRef CAS PubMed.
  92. B. P. Biswal, S. Chandra, S. Kandambeth, B. Lukose, T. Heine and R. Banerjee, J. Am. Chem. Soc., 2013, 135, 5328–5331 CrossRef CAS PubMed.
  93. S. Karak, S. Kandambeth, B. P. Biswal, H. S. Sasmal, S. Kumar, P. Pachfule and R. Banerjee, J. Am. Chem. Soc., 2017, 139, 1856–1862 CrossRef CAS PubMed.
  94. X. Guan, F. Chen, Q. Fang and S. Qiu, Chem. Soc. Rev., 2020, 49, 1357–1384 RSC.
  95. M. Chen, H. Li, C. Liu, J. Liu, Y. Feng, A. D. H. Wee and B. Zhang, Coord. Chem. Rev., 2021, 435, 213778 CrossRef CAS.
  96. P. Pachfule, A. Acharjya, J. Roeser, T. Langenhahn, M. Schwarze, R. Schomacker, A. Thomas and J. Schmidt, J. Am. Chem. Soc., 2018, 140, 1423–1427 CrossRef CAS PubMed.
  97. H. Yu, J. Zhang, X. Yan, C. Wu, X. Zhu, B. Li, T. Li, Q. Guo, J. Gao, M. Hu and J. Yang, J. Mater. Chem. A, 2022, 10, 11010–11018 RSC.
  98. L. Wang, L. Zhang, B. Lin, Y. Zheng, J. Chen, Y. Zheng, B. Gao, J. Long and Y. Chen, Small, 2021, 17, 2101017 CrossRef CAS PubMed.
  99. M. S. Lohse and T. Bein, Adv. Funct. Mater., 2018, 28, 1705553 CrossRef.
  100. H. Wang, H. Wang, Z. Wang, L. Tang, G. Zeng, P. Xu, M. Chen, T. Xiong, C. Zhou, X. Li, D. Huang, Y. Zhu, Z. Wanga and J. Tang, Chem. Soc. Rev., 2020, 49, 4135–4165 RSC.
  101. H. Li, H. Li, Q. Dai, H. Li and J.-L. Brédas, Adv. Theory Simul., 2018, 1, 1700015 CrossRef.
  102. J. Xie, Z. Fang and H. Wang, Polymers, 2022, 14, 1363 CrossRef CAS PubMed.
  103. X. Zhao, P. Pachfule, S. Li, T. Langenhahn, M. Ye, C. Schlesiger, S. Praetz, J. Schmidt and A. Thomas, J. Am. Chem. Soc., 2019, 141, 6623–6630 CrossRef CAS PubMed.
  104. W. Zhong, R. Sa, L. Li, Y. He, L. Li, J. Bi, Z. Zhuang, Y. Yu and Z. Zou, J. Am. Chem. Soc., 2019, 141, 7615–7621 CrossRef CAS PubMed.
  105. J. Chen, X. Taoa, C. Lia, Y. Mab, L. Taoa, D. Zhenga, J. Zhud, H. Lia, R. Lia and Q. Yanga, Appl. Catal., B, 2020, 262, 118271 CrossRef CAS.
  106. Z. Mi, T. Zhou, W. Weng, J. Unruangsri, K. Hu, W. Yang, C. Wang, K. A. I. Zhang and J. Guo, Angew. Chem., Int. Ed., 2021, 60, 9642–9649 CrossRef CAS PubMed.
  107. J. Thote, H. B. Aiyappa, A. Deshpande, D. D. Diaz, S. Kurungot and R. Banerjee, Chem.–Eur. J., 2014, 20, 15961–15965 CrossRef CAS PubMed.
  108. Y.-H. Yao, J. Li, H. Zhang, H.-L. Tang, L. Fang, G.-D. Niu, X.-J. Sun and F.-M. Zhang, J. Mater. Chem. A, 2020, 8, 8949–8956 RSC.
  109. Y. Hou, C.-X. Cui, E. Zhang, J.-C. Wang, Y. Li, Y. Zhang, Y. Zhang, Q. Wang and J. Jiang, Dalton Trans., 2019, 48, 14989–14995 RSC.
  110. M. Luo, Q. Yang, K. Liu, H. Cao and H. Yan, Chem. Commun., 2019, 55, 5829–5832 RSC.
  111. Y. Wang, Q. Yang, F. Yi, R. Lu, Y. Chen, C. Liu, X. Li, C. Wang and H. Yan, ACS Appl. Mater. Interfaces, 2021, 13, 29916–29925 CrossRef CAS PubMed.
  112. K. Gottschling, G. Savasci, H. V. Gonzalez, S. Schmidt, P. Mauker, T. Banerjee, P. Rovo, C. Ochsenfeld and B. V. Lotsch, J. Am. Chem. Soc., 2020, 142, 12146–12156 CrossRef CAS PubMed.
  113. G. B. Wang, F.-C. Zhu, Q.-Q. Lin, J.-L. Kan, K.-H. Xie, S. Li, Y. Geng and Y.-B. Dong, Chem. Commun., 2021, 57, 4464–4467 RSC.
  114. C. Liu, Y. Xiao, Q. Yang, Y. Wang, R. Lu, Y. Chen, C. Wang and H. Yan, Appl. Surf. Sci., 2021, 537, 148082 CrossRef CAS.
  115. Z. Li, T. Deng, S. Ma, Z. Zhang, G. Wu, J. Wang, Q. Li, H. Xia, S.-W. Yang and X. Liu, J. Am. Chem. Soc., 2023, 145, 8364–8374 CAS.
  116. J. Bi, W. fang, L. Li, J. Wang, S. Liang, Y. He, M. Liu and L. Wu, Macromol. Rapid Commun., 2015, 36, 1799–1805 CrossRef CAS PubMed.
  117. S. Kuecken, A. Acharjya, L. Zhi, M. Schwarze, R. Schomacker and A. Thomas, Chem. Commun., 2017, 53, 5854–5857 RSC.
  118. Z. Cheng, W. Fang, T. Zhao, S. Fang, J. Bi, S. Liang, L. Li, Y. Yu and L. Wu, ACS Appl. Mater. Interfaces, 2018, 10, 41415–41421 CrossRef CAS PubMed.
  119. W. Huang, Q. He, Y. Hu and Y. Li, Angew. Chem., Int. Ed., 2019, 58, 8676–8680 CrossRef CAS PubMed.
  120. L. Gao, Y. Niu, S. Razzaque, B. Tian and S. Jin, ACS Catal., 2019, 9, 9438–9445 CrossRef.
  121. M. A. Fávaro, D. Ditz, J. Yang, S. Bergwinkl, A. C. Ghosh, M. Stammler, C. Lorentz, J. Roeser, E. A. Quadrelli, A. Thomas, R. Palkovits, J. Canivet and F. M. Wisser, ACS Appl. Mater. Interfaces, 2022, 14, 14182–14192 CrossRef PubMed.
  122. C. B. Meier, R. Clowes, E. Berardo, K. E. Jelfs, M. A. Zwijnenburg, R. S. Sprick and A. I. Cooper, Chem. Mater., 2019, 31, 8830–8838 CrossRef CAS PubMed.
  123. K. Wang, L. –M. Yang, X. Wang, L. Guo, G. Cheng, C. Zhang, S. Jin, B. Tan and A. I. Cooper, Angew. Chem., Int. Ed., 2017, 56, 14149–14153 CrossRef CAS PubMed.
  124. M. Liu, Q. Huang, S. Wang, Z. Li, B. Li, S. Jin and B. Tan, Angew. Chem., Int. Ed., 2018, 57, 11968–11972 CrossRef CAS PubMed.
  125. S. Zhang, G. Cheng, L. Guo, N. Wang, B. Tan and S. Jin, Angew. Chem., Int. Ed., 2020, 59, 6007–6014 CrossRef CAS PubMed.
  126. M. Liu, X. Wang, J. Liu, K. Wang, S. Jin and B. Tan, ACS Appl. Mater. Interfaces, 2020, 12, 12774–12782 CrossRef CAS PubMed.
  127. M. Liu, K. Jiang, X. Ding, S. Wang, C. Zhang, J. Liu, Z. Zhan, G. Cheng, B. Li, H. Chen, S. Jin and B. Tan, Adv. Mater., 2019, 31, 1807865 CrossRef PubMed.
  128. X. Hu, Z. Zhan, J. Zhang, I. Hussain and B. Tan, Nat. Commun., 2021, 12, 6596 CrossRef CAS PubMed.
  129. P. Audebert, E. Kroke, C. Posern and S.-H. Le, Chem. Rev., 2021, 121, 2515–2544 CrossRef CAS PubMed.
  130. S. Kumar, V. R. Battula, N. Sharma, S. Samanta, B. Rawata and K. Kailasam, J. Mater. Chem. A, 2022, 10, 14568–14575 RSC.
  131. N. Sharma, S. Kumar, V. R. Battula, A. Kumari, A. Giri, A. Patra and K. Kailasam, Chem.–Eur. J., 2021, 27, 10649–10656 CrossRef CAS PubMed.
  132. D. Chen, W. Chen, G. Zhang, S. Li, W. Chen, G. Xing and L. Chen, ACS Catal., 2022, 12, 616–623 CrossRef CAS.
  133. H. Cheng, H. Lv, J. Cheng, L. Wang, X. Wu and H. Xu, Adv. Mater., 2022, 34, 2107480 CrossRef CAS PubMed.
  134. K. Kailasam, J. Schmidt, H. Bildirir, G. Zhang, S. Blechert, X. Wang and A. Thomas, Macromol. Rapid Commun., 2013, 34, 1008–1013 CrossRef CAS PubMed.
  135. K. Kailasam, M. B. Mesch, L. Mçhlmann, M. Baar, S. Blechert, M. Schwarze, M. Schrçder, R. Schomcker, J. Senker and A. Thomas, Energy Technol., 2016, 4, 744–750 CrossRef CAS.
  136. Y. Zhao, C. Wang, X. Han, Z. Lang, C. Zhao, L. Yin, H. Sun, L. Yan, H. Ren and H. Tan, Adv. Sci., 2022, 9, 2202417 CrossRef CAS PubMed.
  137. V. R. Battula, S. Kumar, D. K. Chauhan, S. Samanta and K. Kailasam, Appl. Catal., B, 2019, 244, 313–319 CrossRef CAS.
  138. S. Kumar, N. Sharma and K. Kailasam, J. Mater. Chem. A, 2018, 6, 21719–21728 RSC.
  139. L. Stegbauer, S. Zech, G. Savasci, T. Banerjee, F. Podjaski, K. Schwinghammer, C. Ochsenfeld and B. V. Lotsch, Adv. Energy Mater., 2018, 8, 1703278 CrossRef.
  140. E. Jin, Z. Lan, Q. Jiang, K. Geng, G. Li, X. Wang and D. Jiang, Chem, 2019, 5, 1632–1647 CAS.
  141. K. Wang, Z. Jia, Y. Bai, X. Wang, S. E. Hodgkiss, L. Chen, S. Y. Chong, X. Wang, H. Yang, Y. Xu, F. Feng, J. W. Ward and A. I. Cooper, J. Am. Chem. Soc., 2020, 142, 11131–11138 CrossRef CAS PubMed.
  142. W. Chen, L. Wang, D. Mo, F. He, Z. Wen, X. Wu, H. Xu and L. Chen, Angew. Chem., Int. Ed., 2020, 59, 16902–16909 CrossRef CAS PubMed.
  143. Z. Zhao, Y. Zheng, C. Wang, S. Zhang, J. Song, Y. Li, S. Ma, P. Cheng, Z. Zhang and Y. Chen, ACS Catal., 2021, 11, 2098–2107 CrossRef CAS.
  144. W. Li, X. Huang, T. Zeng, Y. A. Liu, W. Hu, H. Yang, Y.-B. Zhang and K. Wen, Angew. Chem., Int. Ed., 2021, 60, 1869–1874 CrossRef CAS PubMed.
  145. Y. Li, L. Yang, H. He, L. Sun, H. Wang, X. Fang, Y. Zhao, D. Zheng, Y. Qi, Z. Li and W. Deng, Nat. Commun., 2022, 13, 1355 CrossRef CAS PubMed.
  146. H. Wang, H. Ding, X. Meng and C. Wang, Chin. Chem. Lett., 2016, 27, 1376–1382 CrossRef CAS.
  147. A. M. H. Flores, G. Bengasi, K. Baba and N. D. Boscher, ACS Appl. Energy Mater., 2020, 3, 9848–9855 CrossRef.
  148. J. M. Gonclaves, B. A. Iglesias, P. R. Martins and L. Angnes, Analyst, 2021, 146, 365–381 RSC.
  149. M. Joseph and S. Haridas, Int. J. Hydrogen Energy, 2020, 45, 11954–11975 CrossRef CAS.
  150. Z. Fan, K. Nomura, M. Zhu, X. Li, J. Xue, T. Majima and Y. Osakada, Commun. Chem., 2019, 2, 55 CrossRef.
  151. R. Chen, Y. Wang, Y. Ma, A. Mal, X.-Y. Gao, L. Gao, L. Qiao, X.-B. Li, L.-Z. Wu and C. Wang, Nat. Commun., 2021, 12, 1354 CrossRef CAS PubMed.
  152. S. Kandambeth, A. Mallick, B. Lukose, M. V. Mane, T. Heine and R. Banerjee, J. Am. Chem. Soc., 2012, 134, 19524–19527 CrossRef CAS PubMed.
  153. P. Biswal, H. A. V. Gonzalez, T. Banerjee, L. Grunenberg, G. Savasci, K. Gottschling, J. Nuss, C. Ochsenfeld and B. V. Lotsch, J. Am. Chem. Soc., 2019, 141, 11082–11092 CrossRef PubMed.
  154. T. Zhou, X. Huang, Z. Mi, Y. Zhu, R. Wang, C. Wang and J. Guo, Polym. Chem., 2021, 12, 3250–3256 RSC.
  155. W. Weng and J. Gao, Nat. Commun., 2022, 13, 5768 CrossRef CAS PubMed.
  156. L. Cheng, Q. Xiang, Y. Liao and H. Zhang, Energy Environ. Sci., 2018, 11, 1362–1391 RSC.
  157. H. Dong, X.-B. Meng, X. Zhang, H.-L. Tang, J. W. Liu, J.-H. Wang, J.-Z. Wei, F.-M. Zhang, L.-L. Bai and X.-J. Sun, Chem. Eng. J., 2020, 379, 122342 CrossRef CAS.
  158. Y.-P. Zhang, H.-L. Tang, H. Dong, M.-Y. Gao, C.-C. Li, X.-J. Sun, J.-Z. Wei, Y. Qu, Z.-J. Lib and F.-M. Zhang, J. Mater. Chem. A, 2020, 8, 4334–4340 RSC.
  159. C.-C. Li, M.-Y. Gao, X.-J. Sun, H.-L. Tang, H. Dong and F.-M. Zhang, Appl. Catal., B, 2020, 266, 118586 CrossRef CAS.
  160. M.-Y. Gao, C.-C. Li, H.-L. Tang, X.-J. Sun, H. Dong and F.-M. Zhang, J. Mater. Chem. A, 2019, 7, 20193–20200 RSC.
  161. L.-L. Zheng, D. Wang, S.-L. Wu, X.-H. Jiang, J. Zhang, Q.-J. Xing, J.-P. Zou and S.-L. Luo, J. Mater. Chem. A, 2020, 8, 25425–25430 RSC.
  162. H. Huang, B. Xu, Z. Tan, Q. Jiang, S. Fang, L. Li, J. Bi and L. Wu, J. Alloys Compd., 2020, 833, 155057 CrossRef CAS.
  163. Z. Tan, P. Zhang, Q. Chen, S. Fang, G. Huang, J. Bi and L. Wu, Catal. Sci. Technol., 2021, 11, 1874–1880 RSC.
  164. N. Xu, B. Cai, Q. Li, Y. Liu, J. Tang, K. Wang, B. Xu and Y. Fan, J. Alloys Compd., 2021, 871, 159565 CrossRef CAS.
  165. Y. Peng, M. Zhao, B. Chen, Z. Zhang, Y. Huang, F. Dai, Z. Lai, X. Cui, C. Tan and H. Zhang, Adv. Mater., 2018, 30, 1705454 CrossRef PubMed.
  166. F.-M. Zhang, J.-L. Sheng, Z.-D. Yang, X.-J. Sun, H.-L. Tang, M. Lu, H. Dong, F.-C. Shen, J. Liu and Y.-Q. Lan, Angew. Chem., Int. Ed., 2018, 57, 12106–12110 CrossRef CAS PubMed.
  167. H.-Y. Zhang, Y. Yang, C.-C. Li, H.-L. Tang, F.-M. Zhang, G.-L. Zhang and H. Yan, J. Mater. Chem. A, 2021, 9, 16743–16750 RSC.
  168. J. Low, S. Cao, J. Yu and S. Wageh, Chem. Commun., 2014, 50, 10768–10777 RSC.
  169. Y. Xing, L. Yin, Y. Zhao, Z. Du, H.-Q. Tan, X. Qin, W. Ho, T. Qiu and Y.-G. Li, ACS Appl. Mater. Interfaces, 2020, 12, 51555–51562 CrossRef CAS PubMed.
  170. C. Lin, C. Han, L. Gong, X. Chen, J. Deng, D. Qi, Y. Bian, K. Wang and J. Jiang, Catal. Sci. Technol., 2021, 11, 2616–2621 RSC.
  171. T. Zhou, L. Wang, X. Huang, J. Unruangsri, H. Zhang, R. Wang, Q. Song, Q. Yang, W. Li, C. Wang, K. Takahashi, H. Xu and J. Guo, Nat. Commun., 2021, 12, 3934 CrossRef CAS PubMed.
  172. W. Zheng, C.-S. Tsang, L. Y. S. Lee and K.-Y. Wong, Mater. Today Chem., 2019, 12, 34–60 CrossRef CAS.
  173. C. Altintas, I. Erucar and S. Keskin, CrystEngComm, 2022, 24, 7360–7371 RSC.
  174. Y. Li, G. Wen, J. Li, Q. Li, H. Zhang, B. Tao and J. Zhang, Chem. Commun., 2022, 58, 11488–11506 RSC.
  175. R. Kumar, S. N. Ansari, R. Deka, P. Kumar, M. Saraf and S. M. Mobin, Chem.–Eur. J., 2021, 27, 1–31 CrossRef.
  176. D. Shang, D. Li, B. Chen, B. Luo, Y. Haung and W. Shi, ACS Sustainable Chem. Eng., 2021, 9, 14238–14248 CrossRef CAS.
  177. A. Xiong, G. Ma, K. Maeda, T. Takata, T. Hisatomi, T. Setoyama, J. Kubota and K. Domen, Catal. Sci. Technol., 2014, 4, 325–328 RSC.
  178. B. Mishra, A. Alam, B. Kumbhakar, D. D. Díaz and P. Pachfule, Cryst. Growth Des., 2023, 23(6), 4701–4719 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2023