Emerging heterostructured C3N4 photocatalysts for photocatalytic environmental pollutant elimination and sterilization

Yang Ding a, Chunhua Wang b, Lang Pei a, Soumyajit Maitra c, Qinan Mao a, Runtian Zheng d, Meijiao Liu e, Yun Hau Ng b, Jiasong Zhong *a, Li-Hua Chen *f and Bao-Lian Su *df
aCenter of Advanced Optoelectronic Materials, College of Materials and Environmental Engineering, Hangzhou Dianzi University, Hangzhou 310018, China. E-mail: jiasongzhong@hdu.edu.cn
bSchool of Energy and Environment, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong, China
cDepartment of Materials, University of Oxford, Oxford, OX1 3PH, UK
dLaboratory of Inorganic Materials Chemistry (CMI), University of Namur, 61 rue de Bruxelles, B-5000, Namur, Belgium
eDepartment of Chemistry, Zhejiang Sci-Tech University, Hangzhou 310018, China
fState Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, 122, Luoshi Road, 430070 Wuhan, Hubei, China. E-mail: chenlihua@whut.edu.cn; bao-lian.su@unamur.be

Received 8th April 2023 , Accepted 12th May 2023

First published on 17th May 2023


Abstract

Photocatalysis is deemed a highly prominent technology to solve environmental problems such as pollution, CO2 emission and bacterial contamination. As an important photocatalyst, g-C3N4 has attracted a great amount of attention in environmental remediation owing to its good stability, excellent light response, low cost and environmentally friendly properties. However, the pristine g-C3N4 photocatalyst generally suffers from serious photoinduced charge carrier recombination, poor surface active sites, and insufficient visible light harvesting, thereby leading to unsatisfactory photocatalytic performance. Heterostructured C3N4 photocatalysts have recently become a research focus in environmental fields thanks to their fast photoexcited electron–hole pair dissociation, broadened visible light response range, and sufficient photoredox capability. Herein, we critically review the up-to-date developments of heterostructured C3N4 photocatalysts in organic pollutant elimination, heavy metal ion reduction, CO2 conversion and bacterial inactivation. Meanwhile, the strategies for constructing efficient C3N4 based heterostructures with enhanced environmental photocatalytic capability are thoroughly described, which should help readers to quickly acquire in-depth knowledge and to inspire new concepts in heterostructure engineering. Finally, the challenges and opportunities in fabricating heterostructured C3N4 photocatalysts for large-scale and commercial applications are discussed to give a clear study direction in this field.


1. Introduction

Since the pioneering work regarding solar-light-driven water decomposition was reported in 1972,1 the transformation of solar energy into chemical energy has attracted tremendous attention in environmental remediation and sustainable energy evolution.2–6 Furthermore, the photocatalysis method has been considered as a feasible means for harvesting light energy in chemical bonds through different photoexcited reactions, like organic pollutant oxidation, CO2 reduction, H2O splitting, H2O2 preparation and valued organics synthesis.7–15 Over recent decades, a considerable number of semiconductors such as metal oxides, metal sulfides, and nitride materials have been exploited as photocatalysts for various photocatalytic reactions.16–20 Among them, graphitic carbon nitride (g-C3N4) possessing outstanding light-response ability has shown great promise because of its facile synthesis, tunable band gap structure, good chemical stability, and environmentally friendly properties.21–25 Since the pioneering research regarding photoinduced water splitting over g-C3N4 was reported in 2009,26 lots of studies have been conducted to broaden the application of C3N4 based photocatalysts in different photocatalytic reactions such as toxic organic pollutant elimination, heavy metal ion reduction, CO2 conversion and bacterial inactivation.27–30 However, pristine g-C3N4 generally suffers from limited visible light harvesting, sluggish photoexcited electron–hole pair separation and insufficient photoredox potentials, thereby leading to poor photocatalytic performance. Up to now, a variety of modification methods such as ion doping, defect introduction, heterostructure construction, surface modification, and noble metal nanoparticle decoration have been developed for improving photocatalytic reaction rates.31–35 In particular, the combination of g-C3N4 with other semiconductors to obtain heterostructures has proven the most effective. Usually, engineering heterojunctions can effectively increase the visible light response range and promote photoinduced charge dissociation of photocatalysts.36–38 In addition, the oxidation and reduction potentials of the heterostructured photocatalysts with Z-scheme configurations are strengthened, thus improving the photoredox ability and broadening the photocatalytic applications in environmental remediation fields.39–43 More significantly, in comparison to other common photocatalysts like TiO2 and ZnO, bulk C3N4 is much easier to exfoliate into two-dimensional (2D) nanoplates or nanosheets via facilely breaking the van der Waals forces and hydrogen bonds between the layered C3N4 units, which are flexible and able to be coupled with other semiconductor photocatalysts. These features make C3N4 more favorable to be used in establishing heterojunction photocatalysts compared with other semiconductor substrates.44,45

As is well-known, environmental issues such as water pollution, air deterioration and microbial contamination have posed great threats to human beings. Although a variety of methods including adsorption, chlorination, and ultraviolet (UV) irradiation have been adopted for environmental remediation, there are still some drawbacks such as complicated operation and huge energy consumption. Therefore, it is urgent and meaningful to explore a facile, inexpensive, and environmentally friendly strategy to resolve environmental problems. Photocatalytic techniques have been verified to possess significant capability to eliminate organic/inorganic pollutants, purify air and kill bacteria in water via a green and sustainable process.46 In the visible light driven organic pollutant degradation reactions, the photoexcited electrons on the conduction band (CB) of the semiconductor can reduce O2 to form oxidative ˙O2 species, while the photoinduced holes on the valence band (VB) of the semiconductor can combine with OH to produce oxidative ˙OH radicals for further organic substance decomposition or to directly oxidize organic molecules. As for heavy metal ion pollutants like Cr6+, converting Cr6+ into the less toxic Cr3+ can be realized via a reduction process by the photoexcited electrons on the CB of the semiconductors. In addition to pollutant elimination, the reactive radicals like holes, ˙O2, and ˙OH can also exert important functions in microbial inactivation. Therefore, designing semiconductor photocatalysts with fast photoinduced charge separation, favorable generation of reactive species and sufficient photoredox potentials is highly anticipated for efficient photocatalytic environmental applications.47–52

As important visible light response photocatalysts, heterostructured C3N4 photocatalysts with reliable activity in environmental remediation and protection including organic pollutant degradation, heavy metal ion reduction, CO2 conversion, sewage purification and microbial inactivation have been attracting more and more attention.53–56 Some reviews regarding the applications of heterostructured C3N4 photocatalysts in the photocatalytic field have been published.53,56 Nonetheless, a critical review that thoroughly discloses the relationship between the construction of heterostructured C3N4 materials and improved performance in environmental applications is still imperatively required, and fundamentally crucial for further exploration of advanced heterostructured photocatalysts and expanding their application. In this review, the categories and photocatalytic environmental applications, including organic pollutant degradation, heavy metal ion reduction, CO2 conversion, and microbial inactivation, of heterostructured C3N4 photocatalysts are thoroughly outlined and discussed (Fig. 1). Meanwhile, the relationships between the fabrication of efficient C3N4 based heterostructures and improved performance in photocatalytic environmental applications are correlated in depth. Finally, the challenges and perspectives of heterostructured C3N4 photocatalysts in real applications are discussed to offer a clear study direction in heterostructured photocatalyst design and application.


image file: d3qi00657c-f1.tif
Fig. 1 Applications of heterostructured C3N4 photocatalysts.

2. Categories of heterostructured C3N4 photocatalysts

Until now, various semiconductors including metal oxides, metal chalcogenides and Bi based materials have been coupled with g-C3N4 to construct heterojunctions with efficient photocatalytic performance for different applications. According to the mechanism of photoinduced carrier separation and transfer, four main kinds of g-C3N4 based heterostructured photocatalysts in the environmental photocatalysis field are outlined and discussed in this review as follows: type II heterojunction, Z-scheme heterojunction, p–n heterojunction, and cascade electronic band heterojunction (Fig. 2).57–62
image file: d3qi00657c-f2.tif
Fig. 2 Common types of g-C3N4 based heterojunction for photocatalytic environmental applications: (a) type II heterojunction, (b) Z-scheme heterojunction, (c) p–n heterojunction and (d) cascade electronic band heterojunction.

2.1. g-C3N4-based conventional type II heterojunction systems

With regard to the pure C3N4 photocatalyst, the photoinduced electrons on the conduction band (CB) are inclined to come back to the valence band (VB), and the unfavorable recombination of photoexcited electrons and holes occurs, thus leading to the unsatisfactory photocatalytic performance. When a semiconductor with an appropriate electronic bandgap is decorated on the C3N4 photocatalyst, the facile separation of photoexcited charge pairs can be achieved via the interface between the two semiconductors for enhanced photocatalytic activity. As presented in Fig. 2a, the type II g-C3N4 based heterojunction photocatalyst gives rise to convenient photo-generated charge dissociation owing to the staggered bandgap configuration. Fig. 2a (left) depicts a heterojunction that consists of C3N4 and a semiconductor possessing a more positive VB level. Under light irradiation, the photoelectrons can be shifted from the CB of g-C3N4 to the CB of another semiconductor whereas the photoinduced holes are moved from the VB of another semiconductor to the VB of g-C3N4. The above charge transfer mechanism is frequently observed in lots of type II g-C3N4 based heterojunction photocatalysts such as TiO2/g-C3N4, WO3/g-C3N4, and ZnO/g-C3N4.57–59 In addition to the aforementioned staggered bandgap structure, another interlaced bandgap configuration was also revealed in type II g-C3N4 based heterojunction photocatalysts (Fig. 2a, right). In this case, a semiconductor has a more negative CB position compared with that of g-C3N4, and the photoexcited electrons remain on the CB of g-C3N4 while the photoinduced holes remain on the VB of the other semiconductor after photogenerated charge dissociation and migration. In comparison with the previous type II heterojunction (Fig. 2a, left), the latter type II heterojunction is less reported because of the sufficiently positive VB level of g-C3N4 (about +1.9 eV vs. normal hydrogen electrode (NHE)). Although sufficient carrier separation can be reached in type II heterojunction systems, electronic repulsion phenomena may occur and thus limit the photocatalytic performance. Meanwhile, the energy utilization efficiency needs to be further improved. These shortcomings should be addressed by scientists in future research.

2.2. g-C3N4-based Z-scheme heterojunction systems

Although fast photocharge separation and transfer can be achieved in type II g-C3N4 based heterojunction photocatalysts, the insufficient photo-reduction and photo-oxidation energy levels after the two semiconductors couple limit the applications of g-C3N4 based heterojunction photocatalysts. Achieving efficient photoinduced carrier separation while simultaneously retaining sufficient photoredox performance is therefore significantly meaningful for fabricating g-C3N4 based heterojunction photocatalysts. In this context, the Z-scheme g-C3N4 heterojunction systems have been revealed by scientists. Actually, the most optimal means for charge diffusion and migration was found in the photosynthesis procedure of green plants.60 Under sunlight illumination, the charges shift in the plant rootstock via a different route from the type II heterojunction with the assistance of electron mediators. This category of charge migration is the Z-scheme, which offers fast photoexcited hole–electron pair dissociation and simultaneous competent photoredox abilities. Inspired by the natural photosynthesis process, engineering Z-scheme g-C3N4 based heterojunctions has been considered a highly feasible strategy for achieving efficient photocatalytic activity. Fig. 2c (left) displays the typical Z-scheme charge migration route between two semiconductors. It is clear that the photoelectrons from the CB of semiconductor I are shifted to the VB of g-C3N4. Consequently, the photoexcited hole–electron pairs from semiconductor I and g-C3N4 can be effectively dissociated. More importantly, the electrons on the CB of g-C3N4 and holes on the VB of semiconductor I can endow highly sufficient reduction and oxidation abilities, respectively. In order to further improve the photocatalytic ability, a Z-scheme heterostructure with an electron mediator between the two semiconductors has been exploited. As presented in Fig. 2b (right), an electron mediator located at the interface between semiconductor I and g-C3N4 can promote charge migration for further improved photocatalytic performance. In general, noble nanoparticles with good electron conductivity such as Pd and Pt are frequently adopted as electron mediators. In addition to the fantastic Z-scheme charge migration route, the formation of an internal built-in electric field between the two semiconductors is another advantage for promoting carrier dissociation. In some cases, the Z-scheme heterojunction is also denoted as an S-scheme heterojunction.40,41 Despite the different terms, the charge transfer and separation mechanisms for the Z-scheme and S-scheme heterojunctions are almost the same. Recently, g-C3N4 based Z-scheme heterostructures have been classified into three kinds by researchers, the redox mediator solution-phase Z-scheme, solid-state Z-scheme, and direct Z-scheme, which can certainly offer important knowledge for designing heterostructures.44,45

2.3. g-C3N4-based p–n heterojunctions

Apart from the above-mentioned type II and Z-scheme heterostructure photocatalysts, building an inner electric field via engineering a p–n heterojunction between two semiconductors has also been investigated as a feasible means for efficient hole–electron pair separation.61 Commonly, g-C3N4 displays an n-type semiconductor character because of the existence of the electron donors of –NH and NH2 on its surface. As shown in Fig. 2c (left), the Fermi level (EF,n) of n-type g-C3N4 is situated near its CB, whereas the Fermi level (EF,p) of a p-type semiconductor is close to its VB. When g-C3N4 and a p-type semiconductor are combined together, the electrons from g-C3N4 can move to the p-type semiconductor via the interface because of the offset of Fermi levels. As a result, the interface with g-C3N4 becomes positive while the edge of the p-type semiconductor becomes negative. Therefore, an inner electric field is formed between g-C3N4 and the p-type semiconductor. Under light illumination, the photoelectrons transfer from the CB of the p-type semiconductor to the CB of g-C3N4 whereas the photoinduced holes migrate from the VB of g-C3N4 to the VB of the p-type semiconductor due to the driving force from the formed inner electric field, thus achieving fast photogenerated carrier separation and shift, and favorable for increased photocatalytic activity (Fig. 2c, right). Just like the g-C3N4 based p–n heterojunction, g-C3N4 can also form an n–n type heterojunction such as N,S-TiO2/g-C3N424 with increased photocatalytic performance, which can be produced via combining two kinds of p-type semiconductors. After the combination, the n–n type heterojunction presents an inner electric field similar to that of the g-C3N4 based p–n heterojunction for promoted carrier separation. Furthermore, with p–n/n–n heterojunctions, the loading of a co-catalyst has been verified as an efficient way to improve photocatalytic ability.23,24 Ti3C2 serving as co-catalyst combined with different photocatalysts has been reported. Specifically, the low Fermi level of Ti3C2 can promote the generation of a Schottky junction with the semiconductors to function as an electron trapper and facilitate charge dissociation. For example, a Ti3C2/N,S-doped TiO2/g-C3N4 composite was prepared with Ti3C2 acting as co-catalyst to improve charge separation because of the generation of an effective charge transfer pathway involving an n–n heterojunction and Schottky junction.24

2.4. g-C3N4-based cascade electronic band heterojunctions

By coupling a semiconductor with g-C3N4, various heterostructures such as type II and Z-scheme types have been obtained with improved photocatalytic ability. Nevertheless, the shortcoming that the photoinduced charges could still undergo recombination during the migration process to the surface active sites for photoredox reactions is rarely addressed. The photocatalytic ability of heterostructured photocatalysts has still not been fully explored.62 On account of this phenomenon, further combining a proper photocatalyst (semiconductor 2) with rich surface active sites with the heterostructured photocatalyst g-C3N4/semiconductor 1 to fabricate a ternary g-C3N4/semiconductor 1/semiconductor 2 system has been recently developed. As shown in Fig. 2d, when the typical ternary g-C3N4 based heterostructure photocatalyst is excited under light, all three components can generate holes and electrons, and the photoelectrons can shift from g-C3N4 to semiconductor 1 and semiconductor 2 via a cascade transport route whereas the holes move along an opposite direction. This g-C3N4 based cascade electronic band heterojunction gives rise to the quick dissociation of photoinduced carriers and therefore improves the photocatalytic performance.

Overall, a variety of heterojunctions have been designed for enhanced photocatalytic performance, which can clearly suggest directions for design principles of heterostructured C3N4 photocatalysts for efficient environmental applications. Firstly, a heterostructure with facilitated carrier dissociation ability should be the prerequisite. Secondly, for environmental pollutant elimination applications such as organics decomposition and Cr(VI) reduction, sufficient photoredox potentials, which can provide enough driving force to promote pollutant molecule elimination, are of significant importance and should be paid more attention. Lastly, for the preparation strategy of heterostructured C3N4 photocatalysts, the wet-chemical route may lead to a uniform morphology and avoid the undesirable agglomeration, thus ensuring more rich surface reaction sites.

3. Photocatalytic pollutant removal by g-C3N4 based heterostructures

So far, a considerable number of pollutants including dyes, pesticides, heavy metal ions, volatile organic compounds (VOCs), and bacteria have been emitted into air, water, and soil, leading to huge threats to human beings and the environment.63–65 In order to eliminate these hazardous substances, a variety of techniques and devices, such as adsorption, filtration, UV light irradiation and biological decomposition, have been developed and desirable achievements have been realized.65 Nevertheless, these techniques still have some intrinsic shortcomings including being time-consuming and involving complicated operation, the formation of toxic byproducts, and huge energy waste. From a long-term perspective with the strategic goal of carbon peaking and carbon neutralization, exploring a facile, green, sustainable and energy-saving strategy for removing environmental pollutants is becoming a mainstream focus of scientific research. Accordingly, the environmental photocatalysis technique has been attracting much attention owing to its obvious merits in comparison to the traditional techniques66–70 (Fig. 3). In this section, various environmental photocatalysis applications of g-C3N4 based heterostructures including dye decomposition, VOC elimination, heavy metal ion reduction, and bacterial inactivation are thoroughly outlined and discussed. Meanwhile, the related mechanisms for enhanced environmental photocatalysis are elaborated.
image file: d3qi00657c-f3.tif
Fig. 3 Schematic mechanism of environmental photocatalysis technique for pollutant treatment.

3.1. Dye pollutant degradation

Among the various photocatalytic pollutant removal methods, photocatalytic dye (e.g., Rhodamine B (RhB), methylene blue (MB), and methyl orange (MO)) degradation is an effective technique and has been inspiring a great number of studies.71,72 In general, dye pollutant degradation over g-C3N4 based heterostructure photocatalysts comprises the following four major steps to totally mineralize organic molecules. Firstly, organic molecules transfer from the pollutant solution to the surface of the photocatalyst. Secondly, electron–hole pairs can be formed upon sunlight irradiation. Thirdly, photo-oxidation decomposition of organic molecules occurs on the surface of the photocatalyst with the participation of photoinduced oxidative species such as holes, hydroxyl radicals and super oxygen. Finally, the decomposition products like CO2 and H2O are released to the pollutant solution.73–75 Taking advantage of their excellent optoelectronic characteristics and optimized energy level configuration, facile separation and shift of charges, enhanced surface photooxidation rates, and favorable generation of reactive species are achieved on g-C3N4 based heterostructure photocatalysts for organic pollutant degradation and mineralization.76–82

Typically, Li et al. prepared MoS2 nanosheets attached to g-C3N4 (MoS2/C3N4) as a heterostructure photocatalyst through a simple ultrasonic chemical technique for efficient photocatalytic dye decomposition.76 After MoS2 loading, rapid dissociation of photoinduced carriers, broadened visible light absorption and increased photocurrent density were achieved on the MoS2/C3N4 heterostructure. On account of the above merits, the optimal MoS2/C3N4 heterostructure with 0.05 wt% MoS2 loading exhibits the largest photocatalytic RhB degradation rate constant of 0.301 min−1, 3.6 times that of pure C3N4 (Fig. 4a and b). The type II charge transfer pathway in MoS2/C3N4 for photocatalytic RhB degradation was proposed. Specifically, MoS2 acting as an electron trapper can prolong the lifetime of electron charges, and the accumulated holes on the VB of C3N4 directly oxidize RhB molecules (Fig. 4c). Similarly, a type II 3D g-C3N4/TiO2 heterojunction photocatalyst was synthesized for efficient MB degradation.77 The MB degradation rate over 3D g-C3N4/TiO2 is about 4 times that of bulk g-C3N4. In addition to the benefit of the heterojunction in promoting electron–hole pair separation (Fig. 4d), the 3D structure can also elevate the adsorption enrichment ability and endow convenient mass and charge migration channels (Fig. 4e), therefore improving the MB degradation capability. Hao et al. synthesized a macro/mesoporous g-C3N4/TiO2 heterojunction photocatalyst (Fig. 4f and g) via a template free route.78 Owing to the advantages of morphology, porous structure and type II heterostructure, the macro/mesoporous g-C3N4/TiO2 photocatalyst achieved an RhB degradation rate as high as 0.0478 min−1, which is about 7.2 and 3.1 fold that of the pure TiO2 and g-C3N4 photocatalysts, respectively (Fig. 4h).


image file: d3qi00657c-f4.tif
Fig. 4 (a) Time dependent photocatalytic degradation of RhB over the as-obtained photocatalysts and (b) the corresponding degradation rates. (c) Schematic illustration of photoinduced charge transfer over the MoS2/C3N4 heterostructure. Reproduced with permission.76 Copyright 2014, American Chemical Society. (d) Photocatalytic mechanism for organic pollutant elimination over 3D g-C3N4/TiO2. (e) SEM image of 3D g-C3N4/TiO2. Reproduced with permission.77 Copyright 2019, Elsevier. (f and g) SEM pictures of the macro/mesoporous g-C3N4/TiO2. (h) Photocatalytic RhB degradation rates of the as-obtained samples. Reproduced with permission.78 Copyright 2016, Elsevier.

Compared with the type II heterostructure, the Z-scheme heterostructure possesses higher capability for dye removal. In addition to facilitating photoinduced charge separation, the Z-scheme configuration enables sufficient photoredox ability, enabling dye oxidation and mineralization. For example, a Z-scheme 2D/2D g-C3N4/MnO2 heterostructured photocatalyst was prepared through in situ loading MnO2 nanosheets onto g-C3N4 nanosheets via a wet chemical strategy (Fig. 5a).79 As expected, the g-C3N4/MnO2 heterostructured photocatalyst displayed increased photocatalytic RhB degradation ability compared with the pure g-C3N4 or MnO2 sample. The charge transfer mechanism for photocatalytic RhB degradation is depicted in Fig. 5b, wherein the photoexcited electrons on the CB of MnO2 are coupled with the holes on the VB of g-C3N4, giving rise to efficient charge carrier dissociation and visible light harvesting. Moreover, more negative CB and positive VB potentials are obtained after the Z-scheme charge separation and migration, thus delivering enough photoredox ability. Specifically, the holes on the VB of MnO2 can oxidize H2O to form oxidative ˙OH radicals and the photoinduced electrons on the CB of g-C3N4 produce ˙O2 radicals. Both the reactive radicals are responsible for RhB molecule decomposition.


image file: d3qi00657c-f5.tif
Fig. 5 (a) Schematic process for the preparation of a 2D/2D g-C3N4/MnO2 heterostructure and (b) photoexcited charge transfer pathway over the g-C3N4/MnO2 heterostructure. Reproduced with permission.79 Copyright 2018, American Chemical Society. (c) Schematic illustration of the synthesis of a ternary g-C3N4/Al2O3/ZnO photocatalyst. EPR characterization of (d) DMPO-˙O2 and (e) DMPO-˙OH adducts over the as-synthesized photocatalysts. (f) Schematic cascade electron transfer route for MO degradation. Reproduced with permission.81 Copyright 2017, Elsevier.

In addition to binary C3N4 based composites with type II or Z-scheme heterostructures, the ternary C3N4 based heterostructure with a cascade electron energy structure has also aroused great interest owing to its unique and rapid electron shift capability. For instance, a ternary H2SrTa2O7/g-C3N4/Ag3PO4 composite was prepared by a facile impregnation and ion exchange strategy for enhanced photodecomposition of methyl orange (MO).80 Firstly, a binary H2SrTa2O7/g-C3N4 hybrid with an optimal g-C3N4 loading mass ratio of 60 wt% was synthesized. Then, Ag3PO4 is decorated on H2SrTa2O7/g-C3N4 to create the ternary H2SrTa2O7/g-C3N4/Ag3PO4 photocatalyst with further improved photoactivity. An increased interface interaction between H2SrTa2O7, g-C3N4 and Ag3PO4 is realized, which is favorable for separation of photoexcited charge carriers. Accordingly, the ternary H2SrTa2O7/g-C3N4/Ag3PO4 composite exhibits about 45, 44 and 38 fold higher rates of photocatalytic MO removal than those of the H2SrTa2O7, g-C3N4 and H2SrTa2O7/g-C3N4 photocatalysts, respectively. Electron spin resonance measurements indicated that photoinduced holes and superoxide radicals exert a significant function in decomposing MO molecules. Moreover, a cascade electron transfer route is followed on the ternary photocatalyst. Specifically, under light irradiation, the photoinduced electrons migrate from g-C3N4 to H2SrTa2O7 and Ag3PO4, and combine with O2 to generate oxidative superoxide radicals for phenol decomposition while the photogenerated holes on the VB of g-C3N4 directly take part in phenol degradation. This stepwise migration leads to the facile separation of photogenerated carriers and therefore elevates the photocatalytic activity.

Similarly, a ternary g-C3N4/Al2O3/ZnO photocatalyst was obtained through a hydrothermal and calcination route (Fig. 5c).81 The amorphous Al2O3 component with a disorganized atomic arrangement and defects can achieve a more favorable electron acceptance ability from g-C3N4 than that of other crystalline materials. EPR characterization disclosed that the ternary g-C3N4/Al2O3/ZnO photocatalyst exhibits greater ability in producing reactive oxygen species such as superoxide radicals and hydroxyl radicals than that of the binary g-C3N4/ZnO and g-C3N4/Al2O3 samples (Fig. 5d and e), which is favorable for photo-oxidation of dye molecules. Under visible light irradiation, the ternary g-C3N4/Al2O3/ZnO photocatalyst shows about a 1.7 times higher rate for methyl orange degradation than that of pure g-C3N4 due to the prominent cascade electron transfer route. As shown in Fig. 5f, the photoexcited electrons on the CB of g-C3N4 move to the CB of amorphous Al2O3 and then shift to the CB of ZnO via the cascade transport route, which effectively facilitates the hole–electron pair dissociation. Finally, the electrons on the CB of ZnO can combine and activate O2 to produce ˙O2 and ˙OH for oxidizing and mineralizing MO molecules.

On the whole, via constructing a series of heterostructures including the type II heterostructure, Z-scheme and tandem type, fast charge separation and efficient photocatalytic dye decomposition can be realized. However, some aspects still should be clarified. For example, the TiO2/g-C3N4 heterostructure system established via different methods usually displays various charge transfer pathways in dye elimination. The inherent mechanism leading to such a variation for the same system should be revealed by advanced characterization techniques such as in situ XPS.

3.2. VOC elimination

Volatile organic compounds (VOCs) are highly detrimental pollutants in the atmosphere and lead to decreased outdoor and indoor air quality.83–85 Hence, effectively eliminating VOCs has aroused a great deal of concern in pollutant treatment. A variety of strategies such as adsorption, combustion and biological degradation have been adopted for eliminating VOCs. With the fast development of economy and human society, more facile, green and low-cost methods for VOC removal haves attracted great attention.86 Accordingly, photocatalytic oxidation as a clean and sustainable technique has been widely studied for the oxidation decomposition of VOCs. The significant step for photocatalytic oxidation is to form photoinduced hole–electron pairs and dissociate the pairs to achieve photooxidation reaction under light illumination. Specifically, the photoinduced holes with strong oxidative ability can directly decompose VOC molecules. Meanwhile, other reactive oxygen species like ˙O2 and ˙OH can be also generated during the photocatalytic process, which are also useful for VOC molecule elimination.87,88

Typically, a g-C3N4@Bi2WO6 core–shell structured heterostructure was fabricated by in situ C3N4 precursor encapsulation and reassembly.89 As illustrated in Fig. 6a, the precursor molecules were polymerized to generate a thin g-C3N4 layer on the surface of Bi2WO6 nanosheets. The obtained g-C3N4@Bi2WO6 photocatalyst with a g-C3N4 layer thickness of about 1 nm (Fig. 6b and c) displayed the highest visible light driven photocatalytic phenol degradation rate, being about 6 and 2 times that of bulk g-C3N4 and pure Bi2WO6 nanosheets, respectively. Electron spin resonance spectroscopy indicated that the superoxide radicals and hydroxyl radicals are more likely to be generated on g-C3N4@Bi2WO6 compared with pure Bi2WO6 and bulk g-C3N4, which demonstrates that the formation of a core–shell structured heterostructure can enable a high oxidation ability. Accordingly, a type II heterojunction for photoexcited charge transfer on g-C3N4@Bi2WO6 for photocatalytic phenol degradation is depicted in Fig. 6d. Firstly, hole–electron pairs are formed on the CB and VB of the g-C3N4@Bi2WO6 photocatalyst under visible light illumination. Then, the photoinduced electrons can move easily from the g-C3N4 shell into the Bi2WO6 core because the CB of g-C3N4 is higher than that of Bi2WO6, while the photogenerated holes can shift from Bi2WO6 into g-C3N4. O2 molecules can combine with the electrons on the CB of Bi2WO6 to generate superoxide radicals and the holes on the VB of g-C3N4 display high oxidation ability, resulting in an improvement in phenol decomposition of the g-C3N4@Bi2WO6 photocatalyst.


image file: d3qi00657c-f6.tif
Fig. 6 (a) Schematic diagram of the synthesis of the g-C3N4@Bi2WO6 core–shell structured heterostructure and (b and c) HR-TEM images of the g-C3N4@Bi2WO6 core–shell heterostructure. (d) Photoinduced charge transfer mechanism over the g-C3N4@Bi2WO6 heterostructure. Reproduced with permission.89 Copyright 2018, Elsevier. (e) TEM picture of a ternary BiPO4/TiO2/gC3N4 composite. (f) The cascade electron transfer pathway of the BiPO4/TiO2/gC3N4 composite for phenol degradation. Reproduced with permission.90 Copyright 2016, Elsevier.

A ternary BiPO4/TiO2/gC3N4 heterostructured composite was obtained via a facile impregnation route and clearly exhibited an increased photocatalytic phenol degradation rate.90 TEM images suggested the intimate attachment of the three constituents of BiPO4/TiO2/g-C3N4, which is an important precondition for constructing a ternary heterostructure (Fig. 6e). In comparison to the binary TiO2/BiPO4 and TiO2/g-C3N4 heterostructures, the ternary BiPO4/TiO2/g-C3N4 heterostructure photocatalyst exhibited higher photocatalytic activity due to the optimization of its photoinduced carrier dissociation and shift properties. The radical capturing experiments indicated that the ternary BiPO4/TiO2/g-C3N4 heterostructured composite is more able to generate superoxide radical species compared with ternary photocatalysts and pure TiO2, which are directly responsible for decomposing phenol molecules. A schematic diagram of the proposed cascade electronic band gap in the BiPO4/TiO2/g-C3N4 heterostructure is depicted in Fig. 6f. BiPO4 as well as TiO2, and g-C3N4 can be excited by UV and visible light, respectively, to generate hole–electron pairs. Owing to the staggered band gap configuration, the photoelectrons shift from g-C3N4 to TiO2 and BiPO4, whereas the holes move along the opposite direction to participate in the photooxidation of phenol on the VB of g-C3N4.

Although an increased photocatalytic VOC elimination rate can be reached via different kinds of heterostructures, the total decomposition and mineralization of some specific VOC molecules is still complicated and tedious. Therefore, in-depth studies are necessary in this field. For example, the analysis of intermediates during the degradation of VOC molecules is inevitable. Moreover, the photocatalytic technique must guarantee the non-formation of hazardous substances during the decomposition process.

3.3. Heavy metal ion reduction

Hexavalent chromium ions (Cr(VI)) generally exist in wastewater from chromate preparation, electrolyzing, leather tanning, metallurgy, paper making and other industrial manufacturing. Cr(VI) is facilely absorbed by the human body, thus posing a great threat to human beings via the skin, digestion and mucosa.91–93 The toxicity of Cr(VI) can be significantly decreased as Cr(VI) is reduced to Cr(III). Hence, it is greatly meaningful to remove Cr(VI) from wastewater.94 Up to now, a considerable number of works concerning the photocatalytic reduction of Cr(VI) into Cr(III) have been conducted to facilitate its application in wastewater purification. The obtained Cr(III) product is easily eliminated as solid precipitation waste such as Cr(OH)3.95 For the photocatalytic reduction of Cr(VI), the photoinduced electrons from semiconductor photocatalysts play an important role. Therefore, satisfactory photogenerated hole–electron pair separation ability of a photocatalyst is a precondition for Cr(VI) reduction. Moreover, a more negative CB level for photocatalysts is also significant for promoting Cr(VI) reduction. On account of the above considerations, the construction of heterostructured photocatalysts is deemed an ideal solution.96–101

Typically, g-C3N4/ZnO nanorods were synthesized via a simple hydrothermal route for enhanced photocatalytic reduction of Cr(VI).96 SEM and TEM images suggested that the g-C3N4/ZnO nanorods with a diameter of about 50 nm are well dispersed (Fig. 7a–c). A clear boundary between ZnO and g-C3N4 can be found in the TEM image in Fig. 7d, and the plane spacing ascribed to the ZnO (002) plane is 0.26 nm, demonstrating that a heterojunction formed. A type II heterostructure was engineered between g-C3N4 and ZnO without changing the structure of the single components, which increased the visible light absorption region. As illustrated in Fig. 7e, g-C3N4 can be excited under visible light illumination and photoelectrons on the VB shift to the CB, leaving holes on the VB, while ZnO cannot be irradiated by visible light due to its wide band gap. Because the CB of g-C3N4 is more negative than that of ZnO, the photoinduced electrons of g-C3N4 facilely migrate to the CB of ZnO and participate in the photocatalytic reduction of Cr(VI). The type II electron transfer pathway improves the photoinduced carrier separation rate and enhances photocatalytic ability. Similarly, a Cu-ZrO2@g-C3N4 heterostructure was prepared by loading Cu-doped ZrO2 nanoparticles on the surface of g-C3N4via a hydrothermal route.97 An enhanced Cr(VI) photoreduction was achieved due to its wider solar spectrum response range and facilitated charge carrier separation via a type II heterostructure (Fig. 7f and g).


image file: d3qi00657c-f7.tif
Fig. 7 SEM (a and b) and TEM (c and d) images of the obtained g-C3N4/ZnO heterostructure nanorods and (e) mechanism of photoreduction of Cr(VI) over the g-C3N4/ZnO heterostructure. Reproduced with permission.96 Copyright 2020, Elsevier. (f) UV-visible light absorption over the obtained photocatalysts. (g) Mechanism of photoreduction of the Cu-ZrO2@g-C3N4 heterostructure. Reproduced with permission.97 Copyright 2022, Elsevier.

Apart from engineering a heterostructure, the introduction of effective substrates with good electron mobility is also significantly important for enhanced photocatalytic activity. In this context, a ternary CoS2/g-C3N4-rGO hybrid photocatalyst was fabricated via a facile one-pot solvothermal route.98 The CoS2/g-C3N4-rGO hybrid photocatalyst shows increased utilization of visible light and more facile electron–hole pair dissociation in the heterostructure junction. More importantly, a high specific surface area and abundant exposed reactive sites of CoS2 decorated on rGO are achieved. Ultimately, the ternary CoS2/g-C3N4-rGO hybrid photocatalyst achieves a 99.8% elimination rate for Cr(VI) within 2 hours under visible light illumination.

A B-doped g-C3N4 decorated BiVO4 (BCN-BV) composite with a type II p–n heterojunction was synthesized via loading n-type BiVO4 on the surface of p-type B-doped g-C3N4 for enhanced photoreduction of Cr(VI).99 In comparison with pristine BiVO4 and B-doped g-C3N4, CN-BV exhibits a higher photocatalytic reduction rate of Cr(VI) owing to the formation of a type II p–n heterojunction, which can effectively increase visible light utilization and restrict the recombination of hole–electron pairs. Moreover, the TEM image demonstrated the co-existence of monoclinic and tetragonal phase BiVO4, therefore generating an inner heterojunction for the further improvement of photocatalytic ability (Fig. 8a–c). The schematic mechanism for Cr(VI) reduction over the CN-BV composite is depicted in Fig. 8(d). When BiVO4 is combined with BCN, a type II p–n heterojunction is formed. In order to form an equilibrium in the Fermi level potential between the two semiconductors, the Fermi level of BiVO4 is shifted downward whereas the Fermi level of BCN moves upward. As a result, the CB edge position of BiVO4 is lower than that of BCN. Under visible light excitation, both BVO and BCN can be simultaneously irradiated to generate electron–hole pairs. Along with that, an inner electric field is also formed that enhances the movement of photoexcited charge carriers. Because of the co-existence of both monoclinic and tetragonal phased BiVO4, the photoelectrons from the CB of the tetragonal phase can facilely be shifted to the CB of the monoclinic phase, promoting Cr(VI) reduction reaction (Fig. 8d). Accordingly, the type II p–n heterojunction between BiVO4 and BCN not only facilitates the dissociation of the photoexcited hole–electron pairs but also promotes the Cr(VI) reduction process.


image file: d3qi00657c-f8.tif
Fig. 8 XRD patterns (a), HR-TEM image (b) and SAED pattern (c) of the obtained heterostructure photocatalysts. (d) Mechanism of photoreduction of Cr(VI) over the BCN-BV heterostructure. Reproduced with permission.99 Copyright 2019, American Chemical Society. (e) TEM and (f) HR-TEM images of the obtained Ag@Ag3PO4/g-C3N4/NiFe photocatalyst. (g) Band gap structure and (h) mechanism of photoreduction of Cr(VI) over the Ag@Ag3PO4/g-C3N4/NiFe heterostructure. Reproduced with permission.100 Copyright 2018, American Chemical Society.

In addition to creating binary photocatalysts with type II and p–n heterojunctions for efficient Cr(VI) reduction, the quaternary semiconductor composite with a cascade electron migration pathway has also been explored. For example, a quaternary Ag@Ag3PO4/g-C3N4/NiFe layered double hydroxide (LDH) photocatalyst (Fig. 8e and f) was synthesized via a combination of self-assembly and in situ photoreduction.100 To be specific, p-type Ag3PO4 was electrostatically anchored onto the self-assembled n-type g-C3N4/NiFe (CNLDH) LDH composite. Meanwhile, a small number of Ag+ ions were reduced to metallic Ag nanoparticles via the photoexcited electrons and -OH groups on the surface of LDH upon visible light illumination. Both Ag nanoparticles and Ag3PO4 decorated on the thin layers of g-C3N4/NiFe hybrid enable convenient electron and mass transportation. The surface plasma effect of Ag nanoparticles and oxygen vacancies in the NiFe LDH can effectively promote charge separation and is the major reason for the enhanced photocatalytic activity. The band gap alignments of g-C3N4, NiFe LDH, Ag3PO4, and Ag nanoparticles before and after combination are systematically depicted in Fig. 8g. Before the combination, the band alignments of the four components are messy, and unable to enable the separation and migration of photoinduced carriers. After the combination, a cascade electron migration pathway in the Ag@Ag3PO4/g-C3N4/NiFe LDH was observed, in which the photoelectrons can flow from Ag3PO4 to g-C3N4 and then to NiFe LDH because of the Fermi level alignment of the constituent photocatalyst, which is favorable for the shift of photo-charge for Cr(VI) reduction (Fig. 8h).

In this section, g-C3N4 based heterostructures for dye decomposition, VOC elimination, and heavy metal ion reduction are thoroughly outlined and discussed. Owing to the convenient photoinduced carrier separation and sufficient photoredox potential after the formation of a heterostructure, improved environmental photocatalytic performance is achieved. On the whole, photoinduced holes are responsible for dye and VOC molecule oxidation while the heavy metal ion reduction is majorly determined by the photoelectrons. The experimental details and corresponding performance of various g-C3N4 based heterostructure photocatalysts in organic pollutant elimination and Cr(VI) reduction are listed in Table 1.

Table 1 Experimental details of various g-C3N4 based heterostructure photocatalysts for pollutant treatment
Photocatalyst Conditions Reaction Performance Type of heterojunction Ref.
MoS2/C3N4 Xenon lamp, 0.1 g l−1 RhB degradation 0.301 min−1 Type II 76
3D g-C3N4/TiO2 Xenon lamp, 0.5 g l−1 MB degradation 0.777 h−1 Type II 77
Macro/mesoporous g-C3N4/TiO2 Xenon lamp, 2 g l−1 RhB degradation 0.0478 min−1 Type II 78
g-C3N4/MnO2 Xenon lamp, 1 g l−1 RhB degradation 91.3%, 60 min Z-scheme 79
H2SrTa2O7/g-C3N4/Ag3PO4 λ ≥ 400 nm, 0.5 g l−1 MO degradation 98%, 8 min Cascade electronic band 80
g-C3N4/Al2O3/ZnO Xenon lamp, 1 g l−1 MB degradation 0.0243 min−1 Cascade electronic band 81
CN/rGO@BPQDs Xenon lamp, 1 g l−1 RhB degradation 97%, 20 min n–n type 82
g-C3N4@Bi2WO6 Xenon lamp Phenol degradation 0.078 h−1 Type II 89
BiPO4/TiO2/g-C3N4 Hg–Xe lamp, 1 g l−1 Phenol degradation 7.2 mol min−1 Cascade electronic band 90
g-C3N4/ZnO Xenon lamp, 1 g l−1 Cr(VI) Reduction 98%, 90 min Type II 96
g-C3N4/Cu-doped ZrO2 AM 1.5G, 1 g l−1 Cr(VI) Reduction 0.0094 h−1 Type II 97
CoS2/gC3N4-rGO Xenon lamp, 0.5 g l−1 Cr(VI) Reduction 99.8%, 120 min Type II 98
B-doped g-C3N4/BiVO4 Xenon lamp, 1 g l−1 Cr(VI) Reduction 85%, 30 min p–n type 99
Ag@Ag3PO4/g-C3N4/NiFe LDH Sunlight, 1 g l−1 Cr(VI) Reduction 97%, 120 min Cascade electronic band 100
g-C3N4/UiO-66 Xenon lamp Cr(VI) Reduction 0.1102 min−1 Z-scheme 101


4. CO2 reduction

The greenhouse effect induced by CO2 emission has aroused great concern during the past few decades, as it causes global warming and poses significant challenges to the development of society and economy.102–104 With environmental deterioration and energy shortages, the utilization of fossil fuels has significantly increased, giving rise to a large amount of CO2 emission to the atmosphere and resulting in an environmental crisis. Among the various strategies for alleviating the results of CO2 emission, the adoption of abundant and clean sunlight energy for reducing CO2 into value-added products and fuels is deemed an ideal protocol to simultaneously address both environmental deterioration and energy shortage problems105–108 (Fig. 9). However, the intrinsic inertness and thermodynamic stability of CO2 molecules with a C[double bond, length as m-dash]O bond breaking energy of around 750 kJ mol−1 lead to the requirement of large energy inputs for light driven CO2 conversion.109,110 Up to now, a variety of photocatalysts have been explored for photoreduction of CO2. Nevertheless, several drawbacks such as unsatisfactory energy conversion efficiency, low selectivity, and inability to restrict the competing reaction of hydrogen formation are still present that limit the development of advanced CO2 conversion.111–115 Therefore, the design and fabrication of highly efficient photocatalysts is a core challenge for achieving elevated performance in the photoreduction of CO2.
image file: d3qi00657c-f9.tif
Fig. 9 Schematic diagram of photocatalytic CO2 conversion.

Among a series of semiconductor photocatalysts, heterostructured C3N4 photocatalysts with appropriate electronic structures have been considered as a promising candidate for increased photoreduction of CO2 owing to their long lifetime of photoexcited charge carriers via the spatial dissociation of holes and electrons within the interface region.116–122 As an efficient visible light converter, the CdS semiconductor has an outstanding optical response and has been frequently adopted as a photo-harvester for various photoredox reactions.123–125 In this regard, employing CdS to hybridize with a C3N4 photocatalyst to construct a type II CdS/C3N4 heterostructure with improved CO2 conversion capability was achieved via a simple photoinduced deposition technique.116 The SEM image in Fig. 10a shows that the CdS nanoparticles are tightly decorated on the surface of C3N4. The HR-TEM picture of CdS/C3N4 suggests that the distinct lattice fringes correspond to the well-defined crystal structures of C3N4 and CdS, indicating the formation of a heterostructure (Fig. 10b). A proposed mechanism for reducing CO2 over the CdS/C3N4 heterostructure is depicted in Fig. 10c. Hole–electron pairs are formed on the CdS and C3N4 semiconductors upon visible light illumination. Because CdS has a more negative CB level compared with that of BCN, the generation of an internal electric field occurs, which is located in the space charge position, therefore promoting the dissociation and migration of photoexcited hole–electron pairs. Afterwards, the photoelectrons can shift quickly from the CB of CdS to that of C3N4. The accumulated photoelectrons react with the adsorbed CO2 molecules on the surface of the CdS/C3N4 heterostructure to generate CO. Meanwhile, the holes migrate from the VB of C3N4 to that of CdS and are finally quenched by the sacrificial agent TEOA. Thus, the redox cycle of the light driven CO2 conversion is accomplished. Specifically, the optimal CdS/C3N4 heterostructure displays excellent performance in the photoreduction of CO2 with a CO productivity of about 13 μmol h−1, being 10 fold that of pristine C3N4 (Fig. 10d).


image file: d3qi00657c-f10.tif
Fig. 10 SEM (a) and HR-TEM (b) images of a CdS/C3N4 heterostructure. (c) Mechanism of photoreduction of CO2 over the CdS/C3N4 heterostructure. Reproduced with permission.116 Copyright 2018, American Chemical Society.  (d) Photocatalytic CO and H2 evolution rate over the obtained samples. (e) Photocatalytic CO2 reduction performance of different samples. (f) Mechanism of photoreduction of CO2 over the In2O3@g-C3N4 heterostructure. Reproduced with permission.117 Copyright 2014, Elsevier. SEM (g) and HR-TEM (h) images of the obtained g-C3N4/NaNbO3 nanowire photocatalyst, and the (i) corresponding photoinduced charge transfer for CO2 reduction. Reproduced with permission.118 Copyright 2014, American Chemical Society.

Similarly, Cao et al. reported the in situ growth of In2O3 nanoparticles on the surface of g-C3N4 nanosheets.117 The resulting In2O3@g-C3N4 hybrid structures exhibited a considerable enhancement in photocatalytic CO2 conversion into CH4 (Fig. 10e), resulting from convenient charge separation and transfer via a type II heterostructure between the In2O3 and g-C3N4 components (Fig. 10f). Zou's team fabricated a g-C3N4/NaNbO3 nanowire photocatalyst by loading polymeric g-C3N4 on NaNbO3 nanowires.118 NaNbO3 possesses a unique crystal structure comprising a framework of corner-shared [NbO6] octahedral units, which is beneficial for enhancing charge movement in the crystal. Meanwhile, the NaNbO3 nanowires generally offer higher surface-to-volume ratios and provide more efficient ballistic charge migration along the single nanowire than the diffusive shift in powdered photocatalysts (Fig. 10g and h). Under visible light illumination, the photoinduced charges shift along a type II route (Fig. 10i). On account of the above merits in structure, composition and charge transfer mechanism, the g-C3N4/NaNbO3 heterojunction photocatalyst displays almost 8 times higher CO2 reduction than that of single C3N4 upon visible light irradiation.

Although a large number of works combining C3N4 with both broadband (ZnO and TiO2) and narrowband (In2O3, NaNbO3, Ag3PO4, and SnS2) semiconductors for photocatalytic CO2 reduction were carried out,116–122 achieving high CO2 conversion efficiency and selectivity is still challenging. In the combination of two semiconductors, the Z-scheme electron transfer mechanism has been found to be the optimal strategy for photocatalytic reactions. In this case, the photoexcited electrons with strong reduction capability are situated in one semiconductor, while photoinduced holes with sufficient oxidizing potential are located in the other semiconductor, which can be subsequently used for corresponding surface reactions. Therefore, fast charge separation and efficient photocatalytic CO2 conversion along with a high selectivity of products are reached. Unfortunately, up to now, most of the developed Z-scheme heterostructures for photoreduction of CO2 have employed an additional sacrificial reagent, which is known as the indirect Z-scheme mechanism. Therefore, for large scale photocatalytic CO2 conversion reaction with C3N4, developing a direct Z-scheme heterostructured photocatalyst with an appropriate electronic structure match is urgently required to realize efficient spatial dissociation of charge carriers and thus ideal performance as well as high product selectivity.

Typically, a series of Z-scheme g-C3N4/FeWO4 heterostructure photocatalysts (Fig. 11a) was synthesized for enhanced CO2 reduction performance with high selectivity into CO as solar fuel upon solar light illumination.119 Specifically, the Z-scheme heterostructure with an advanced charge transfer route exhibited an increased CO formation rate of 6 μmol g−1 h−1 at room temperature, which is around 6 and 15 times that of the pure C3N4 and FeWO4 photocatalysts (Fig. 11b). In addition, the product selectivity for CO was 100% over other carbon products from CO2 reduction. As depicted in Fig. 11c, FeWO4 possesses a relatively negative CB position with respect to the VB level of C3N4, which is favorable for achieving the Z-scheme charge shift mechanism. Therefore, the photoelectrons on the CB of C3N4 having sufficient reducing ability are well preserved (which may otherwise recombine with the photoinduced holes from the VB), while the photoexcited holes on the VB of FeWO4 with strong oxidizing capability remain. On account of the above illustration, more photoelectrons are able to take part in the photoreduction of CO2. As a result, both CO production and selectivity are elevated in the C3N4/FeWO4 heterostructure photocatalyst.


image file: d3qi00657c-f11.tif
Fig. 11 (a) HR-TEM image of the g-C3N4/FeWO4 heterostructure. (b) Time dependent photocatalytic CO productivity over the synthesized samples. (c) Mechanism of photoreduction of CO2 and corresponding charge migration route in the g-C3N4/FeWO4 heterostructure. Reproduced with permission.119 Copyright 2019, American Chemical Society. (d) CO2 adsorption isotherms of the prepared g-C3N4, SnS2 and g-C3N4/SnS2 samples. (e) Schematic illustration of photoreduction of CO2 and charge migration pathway of the g-C3N4/SnS2 heterostructure. (f) In situ FTIR spectra of the g-C3N4/SnS2 photocatalyst under different conditions. Reproduced with permission.120 Copyright 2017, Elsevier. (g) UV-visible light absorption spectra of the fabricated photocatalysts. (h) Photocatalytic reduction of CO2 over the fabricated photocatalysts and the (i) corresponding photoinduced charge transfer for CO2 reduction over the ternary Ag/Ag3PO4/g-C3N4. Reproduced with permission.122 Copyright 2015, American Chemical Society.

Similarly, a direct Z-scheme g-C3N4/SnS2 heterojunction was obtained by in situ decorating SnS2 quantum dots on the surface of g-C3N4 by a convenient hydrothermal route.120L-Cysteine as the sulfur precursor can also graft ammine species onto g-C3N4 during the hydrothermal treatment, which significantly promotes the CO2 uptake of the photocatalyst (Fig. 11d). As illustrated in Fig. 11e, Z-scheme charge transfer occurs under light irradiation, with the photoelectrons on SnS2 combining with the photoinduced holes on g-C3N4, which promotes the extraction and utilization of photoelectrons on g-C3N4. As expected, the g-C3N4/SnS2 heterojunction shows superior photocatalytic CO2 reduction in comparison to the individual g-C3N4 and SnS2 samples, which is attributed to the direct Z-scheme charge transfer and enhanced CO2 adsorption ability. In situ FTIR spectra suggested that HCOOH is formed as an intermediate during light driven CO2 reduction (Fig. 11f), which can only be produced by g-C3N4 based on the energy level of the photoexcited electrons, further proving the existence of the Z-scheme heterostructure in the g-C3N4/SnS2 system. A g-C3N4/ZnO composite microsphere was synthesized via an electrostatic self-assembly protocol.121 This method skillfully exploited the opposite surface charge of g-C3N4 and ZnO, acquiring a hierarchical structure with close contact between them. The composite achieves increased light absorption and convenient photoinduced charge transfer originating from the Z-scheme heterostructure. Therefore, an increased photocatalytic CO2 reduction rate was attained. Specifically, g-C3N4/ZnO exhibits a CH3OH productivity of 1.32 μmol h−1 g−1, which is around 2.1 and 4.1 times that of the pure ZnO and g-C3N4, respectively.

In addition to the binary Z-scheme heterojunction discussed above, ternary Z-scheme heterojunctions containing noble metal ions have also been developed for efficient photoreduction of CO2. For example, a ternary Ag/Ag3PO4/g-C3N4 composite photocatalyst was fabricated using an in situ deposition route.122 As shown in Fig. 5, the absorption of Ag3PO4/g-C3N4 composites is progressively enhanced with increasing Ag3PO4 content, which favors the efficient utilization of solar light and elevates the photocatalytic activity (Fig. 11g). Moreover, it can be noted that the ternary composites also display light absorption in the range of 520–700 nm, which can be ascribed to the plasmonic effect of Ag nanoparticles. Accordingly, the ternary photocatalyst exhibits a prominent CO2 reduction rate of 57.5 μmol h−1, which is about 6.1 times that of pristine g-C3N4 (Fig. 11h). The mechanism for the photoreduction of CO2 over the ternary Ag/Ag3PO4/g-C3N4 composite is depicted in Fig. 11i, wherein Ag nanoparticles can act as a charge transportation bridge to activate the Z-scheme Ag3PO4/g-C3N4 system. Because the CB edge of Ag3PO4 is more negative than the Fermi level of the Ag nanoparticles, the photoexcited electrons on the CB of Ag3PO4 thus shift to the Ag nanoparticles. Meanwhile, the photoinduced holes on the VB of g-C3N4 can move to metallic Ag and further combine with the photoelectrons. This type of charge shift pathway efficiently improves the association of hole–electron pairs and enables the electrons and holes to stay on the CB of g-C3N4 and VB of Ag3PO4, respectively, which is favorable for the photoredox capability. Owing to the negative CB position of g-C3N4, these photoelectrons possess strong reduction ability and can facilely reduce CO2 into CO, CH4, CH3OH, and CH3CH2OH.

By effectively separating the photoinduced charges and maintaining appropriate photoreduction capability, the photoreduction of CO2 into organics was achieved. However, the major product of CO2 reduction is CO over most g-C3N4 based heterostructures. Thus, more focused heterostructure engineering should be carried out to facilitate the production of more valuable C1 and C2 products such as methanol, ethylene and ethanol. The experimental details and corresponding performance of various g-C3N4 based heterostructure photocatalysts in CO2 reduction are listed in Table 2.

Table 2 Experimental details of various g-C3N4 based heterostructure photocatalysts for CO2 reduction
Photocatalyst Conditions Reaction Performance Type of heterojunction Ref.
CdS/BCN Xe light, bpy, acetonitrile, TEOA, and CoCl2 CO2 reduction CO productivity of 250 μmol h−1 g−1 Type II 116
In2O3/g-C3N4 Xe light, 0.5 g l−1 CO2 reduction CH4 productivity of 76.7 ppm Type II 117
g-C3N4/NaNbO3 Xe light, photocatalyst (50 mg) CO2 reduction CH4 productivity of 6.4 μmol h−1 g−1 Type II 118
C3N4/FeWO4 Xe light, 2.5 g l−1 0.5 M Na2SO3 CO2 reduction CO productivity of 6 μmol g−1 h−1 Z-scheme 119
g-C3N4/SnS2 Xe light, 5 g l−1 CO2 reduction CH3OH productivity of 2.3 μmol g−1 Z-scheme 120
g-C3N4/ZnO microspheres Xe light, 10 g l−1 0.5 M Na2SO3 CO2 reduction CH3OH productivity of 1.32 μmol h−1 g−1 Z-scheme 121
Ag3PO4/g-C3N4 Xe light, 2.5 g l−1 CO2 reduction CO2 reduction rate of 57.5 μmol h−1 gcat−1 Z-scheme 122


5. Photocatalytic bacteria inactivation over g-C3N4 based heterostructured photocatalysts

Water contamination by bacteria is becoming a tremendous threat to human health.126–128 Although traditional techniques like chlorination and ozonation have proven to be efficacious in the disinfection of water, they suffer from some shortcomings, such as high costs, tedious operation and huge energy consumption.129–132 Photocatalytic bacteria inactivation via a green, facile, and sustainable process has attracted much attention (Fig. 12). The design of highly efficient photocatalysts is actually the key factor for achieving efficient bacteria inactivation.133–135 In fact, owing to their extraordinary photoactivity and stability, g-C3N4 based heterostructures have significant potential in disinfecting water and great progress has been achieved.
image file: d3qi00657c-f12.tif
Fig. 12 Illustrated mechanisms of photocatalytic bacteria inactivation by common semiconductors. Reproduced with permission.132 Copyright 2018, Frontiers Media S.A.

For example, a series of g-C3N4/xAgBr composites was obtained by loading various amounts of AgBr nanoparticles on g-C3N4.136 Compared with the pure g-C3N4 and AgBr samples, the disinfection capability of g-C3N4/xAgBr was evidently improved. As shown in Fig. 13a, g-C3N4/AgBr displays the highest rate of E. coli cell inactivation. The control experiment showed that the cell amount does not change in the absence of photocatalyst or in the dark, suggesting that photocatalyst and light are crucial for bacteria inactivation. With higher AgBr loading, the disinfection activities of g-C3N4-2AgBr and g-C3N4-4AgBr are attenuated because the agglomeration of AgBr nanoparticles on the surface of g-C3N4 restricts photoinduced carrier separation. In order to directly reveal the evolution of E. coli cells during the disinfection procedure, SEM images of E. coli cells were recorded at different inactivation times. As depicted in Fig. 13b, the E. coli cells show a smooth appearance before light irradiation. After visible light photocatalytic disinfection with g-C3N4-AgBr for 0.5 h, the cells become cataplastic (Fig. 13c), exhibiting destroyed cell wall and damaged cytoplasm. After 1 h of visible light illumination (Fig. 13d), the cells appeared to be totally broken down and effective disinfection was realized. Moreover, the scavenger trapping experiments indicated that photogenerated holes have the major role in the bacterial disinfection process (Fig. 13e). Moreover, the disinfection ability of g-C3N4-AgBr toward S. aureus was slower than that of Gram-negative E. coli., which was attributed to the distinct cell wall configuration of the two kinds of cells (Fig. 13f).


image file: d3qi00657c-f13.tif
Fig. 13 (a) Time dependent photocatalytic disinfection efficiencies of E. coli over g-C3N4 and g-C3N4/xAgBr. SEM pictures of E. coli cells at various photocatalytic disinfection times of (b) 0 min, (c) 30 min, and (d) 60 min. (e) Disinfection performance of E. coli over g-C3N4-AgBr with the addition of different scavengers and (f) disinfection ability of S. aureus over g-C3N4 and g-C3N4-AgBr. Reproduced with permission.136 Copyright 2017, Elsevier.

A variety of ZnO/g-C3N4 heterojunction photocatalysts with increased disinfection performance were synthesized by a thermal polycondensation route (Fig. 14a).137 Under simulated sunlight illumination, the ZnO/g-C3N4 heterojunction photocatalyst with 10 wt% ZnO loading displays the optimum bactericidal efficiency in natural water. Specifically, a bactericidal rate of 97.4% is achieved after 60 min of photocatalytic reaction with the addition of 10% ZnO/g-C3N4, while a bactericidal rate of 47.3% was reached for pure g-C3N4 (Fig. 14b). The results indicate that the construction of the ZnO/g-C3N4 heterojunction leads to enhanced photocatalytic disinfection. More in-depth research demonstrated that ZnO/g-C3N4 heterojunction photocatalysts give rise to increased H2O2 production, which can be used for in situ bacterial inactivation in water. As shown in Fig. 14c, after the end of photocatalytic disinfection, the amount of colonies continually reduces. Meanwhile, the concentration of H2O2 gradually decreases. This phenomenon can be ascribed to the fact that there is still a certain amount of H2O2 in the reaction system after photocatalytic disinfection, which can kill and restrict the survival of bacteria. Hence, it was demonstrated that the disinfection mainly relies on the H2O2 in the solution. Moreover, reactive oxygen species including ˙O2 and ˙OH are easily generated on the ZnO/g-C3N4 photocatalyst due to the facile photoinduced hole–electron pair separation and shift via a Z-scheme heterostructure (Fig. 14d), and they are also responsible for the efficient photocatalytic bacterial inactivation.


image file: d3qi00657c-f14.tif
Fig. 14 (a) Preparation process of a ZnO/g-C3N4 heterojunction. (b) Sterilization rates for bacteria over the obtained photocatalysts. (c) Residual sterilization ability after the photocatalytic process. (d) Pathway of H2O2 formation and photocatalytic sterilization over the ZnO/g-C3N4 heterojunction. Reproduced with permission.137 Copyright 2021, Elsevier. (e) HRTEM image of g-C3N4/NiFe2O4. (f) Radical trapping experiments for photocatalytic inactivation of A. flavus over g-C3N4/NiFe2O4 under visible light illumination. (g) A. flavus colony on peanuts after the photocatalytic process in the presence of g-C3N4/NiFe2O4. Reproduced with permission.140 Copyright 2021, Elsevier.

Aspergillus flavus can cause crop production reduction and generate mycotoxin, thus causing huge threats to human beings.138,139 In this regard, a Z-scheme g-C3N4/NiFe2O4 (CN/NFO) heterostructure photocatalyst was prepared via a facile hydrothermal method for efficient photocatalytic Aspergillus flavus inactivation.140 TEM characterization suggested that the ultrathin g-C3N4 nanosheets homogeneously and intimately combine with rhombic NiFe2O4 nanosheets to form a g-C3N4/NiFe2O4 heterostructure (Fig. 14e), which can provide more sufficient surface active sites for photocatalytic processes and promote photoexcited charge separation and transfer, which are beneficial to the photocatalytic ability. Among a series of CN/NFO composites with various amounts of g-C3N4, 0.2CN/NFO presented the highest capability for Aspergillus flavus inactivation with a bactericidal rate of over 90% under visible light illumination for 1.5 h. The efficient photocatalytic disinfection performance was attributed to the optimal photoexcited charge separation, excellent photoelectric properties and appropriate band structure. Radical trapping tests indicated that ˙O2 and ˙OH are major active species for killing Aspergillus flavus (Fig. 14f). In general, peanut is highly vulnerable to Aspergillus flavus contamination. Therefore, the photocatalytic disinfection ability of the 0.2CN/NFO photocatalyst was checked on contaminated peanuts. As shown in Fig. 14g, it is clear that the colony unit number of Aspergillus flavus on the peanuts progressively decreased with prolonged irradiation time.

Although the g-C3N4 based heterojunction photocatalysts have displayed certain promise in the inactivation of several bacteria, the precise optimization of the photocatalytic process is still needed for overall microbial inactivation. In particular, finding competent materials with a sufficient photoredox potential, low cost, and long-term durability are preconditions for the photocatalytic disinfection of microbes. In addition, the facile separation of photocatalyst powders from aqueous solution after the disinfection procedure should be addressed.

6. DFT calculations of heterostructured photocatalysts

Since there are many DFT calculations of the above heterostructured C3N4 photocatalysts, we thus added one section and briefly highlighted the important points and most representative cases of DFT calculations including Fermi level, work function and electron density difference for heterostructured C3N4 photocatalysts.

In order to investigate the heterojunction generated between two different semiconductor photocatalysts, their work functions have frequently been calculated. Generally, electrons will flow from a semiconductor with a smaller work function to another semiconductor with a larger work function via the interfacial heterojunction. Fig. 15a depicts a schematic configuration of a g-C3N4/MnO2 hybrid for calculating the corresponding Fermi levels of MnO2 and g-C3N4. The values of work function for g-C3N4 and MnO2 were determined to be 4.5 and 6.8 eV, respectively (Fig. 15b), implying that the electrons will transfer from g-C3N4 to MnO2via the heterojunction. As a result, g-C3N4 becomes a positive component whereas MnO2 presents a negative feature close to the heterojunction interface. The above conclusion was further verified by the electron density difference diagram displayed in Fig. 15c. The yellow and cyan domains denote electron accumulation and depletion, respectively, demonstrating that electrons flee from g-C3N4 to MnO2via the heterojunction. Similarly, electron migration from g-C3N4 to SnS2via the interface of the g-C3N4/SnS2 heterojunction was verified by the calculated work functions of g-C3N4 and SnS2 (Fig. 15d and e).


image file: d3qi00657c-f15.tif
Fig. 15 (a) Schematic configuration of the g-C3N4/MnO2 hybrid. (b) The calculated work functions of g-C3N4 and MnO2 materials, respectively. (c) Electron density difference diagram of the g-C3N4/MnO2 hybrid. Reproduced with permission.79 Copyright 2018, American Chemical Society. The calculated work functions for (d) g-C3N4 and (e) SnS2. Reproduced with permission.120 Copyright 2017, Elsevier.

7. Conclusions and outlook

g-C3N4 is a highly promising semiconductor photocatalyst as a substitute for the traditional TiO2 owing to its low cost, good stability and abundance. Nevertheless, the unfavorable recombination of photoexcited charges and insufficient visible light utilization capability of pristine g-C3N4 seriously reduce its photocatalytic activity and prohibit its practical application. Among a variety of modification strategies, only the rational fabrication of heterojunctions with two or more semiconductor photocatalysts can combine the merits of multicomponents to facilitate photo-irradiated charge dissociation, improve visible light harvesting and retain the sufficient redox potential of hole–electron pairs simultaneously.

This comprehensive review summarizes the categories of g-C3N4 based heterojunction photocatalysts, including the type-II heterojunction, Z-scheme heterojunction, p–n heterojunction and cascade electronic band heterojunction. The applications of g-C3N4 based heterojunctions in environmental fields involving organic pollutant removal, Cr(VI) ion reduction, CO2 conversion and bacterial disinfection are thoroughly outlined and discussed. Briefly, on account of the strong oxidation ability of the g-C3N4 based heterojunction, efficient organic pollutant removal and bacterial disinfection can be realized, whereas promoted Cr(VI) ion reduction and CO2 conversion over the g-C3N4 based heterojunction result from the sufficient photoreduction potential due to the formation of the heterojunction. Moreover, the construction of a heterostructure also facilitates these environmental applications driven by visible light illumination, thus broadening their practical applications.

Certainly, the design and fabrication of g-C3N4 based heterostructure photocatalysts will still be further developed and expanded in the future. Up to now, great progress and achievements have been made in the utilization of g-C3N4 based heterojunction photocatalysts in environmental fields. Nevertheless, there are still some challenges and shortcomings in this promising field.

First of all, although g-C3N4 based heterojunction photocatalysts actually give rise to increased photocatalytic capability in environmental applications, it should be apparent that the photocatalytic rate is still far from commercial utilization standards and the large-scale production of uniform heterojunction structures is highly challenging. Secondly, the photoexcited charge transfer pathway in various types of g-C3N4 based heterojunctions and the corresponding photocatalytic mechanism should be further elucidated. Therefore, some advanced characterization methods, like in situ analysis technique, in situ XPS, and even theoretical calculations should be well explored and applied. Thirdly, the stability of heterostructured C3N4 photocatalysts for environmental applications should be addressed. The heterostructured C3N4 photocatalysts fabricated via calcination treatment generally show better stability and rigidity than those obtained from wet chemical methods such as hydrothermal reaction in long-term photocatalytic reactions. However, calcination treatment often leads to an aggregated morphology with a low surface area and poor surface active sites. Thus, how to balance good stability and uniform morphology with plentiful surface active sites should be deeply investigated.

Finally, the construction of heterojunction photocatalysts is a typical surface-based process. Therefore, the photoexcited charge transfer will be faster when a larger contact area is present at the heterojunction interface. However, bulk g-C3N4 adopted in lots of g-C3N4 based heterojunctions generally possesses a low surface area and insufficient active sites, limiting the improvement of photocatalytic activity and exploration of exact active sites. The nanostructured g-C3N4 photocatalysts like nanosheets, nanotubes and nanowires possessing a large surface area, fast charge transfer, and sufficient charge redox capability are thus considered to have great potential in g-C3N4 based heterojunctions, and this could be a hopeful protocol to drastically enhance the photocatalytic performance and precisely reveal the exact active sites.

All in all, we cordially hope that this review can offer deep knowledge to readers and spur on the development of some new concepts and thoughts in manufacturing highly efficient g-C3N4 based heterojunction photocatalysts for commercial applications.

Author contributions

Conceptualization, J. Z., L.-H. C. and B.-L. S.; investigation, Y. D.; writing – original draft, Y. D.; writing – review & editing, Y. D., C. W., L. P., S. M., M. Q., R. Z., M. L., Y. N., J. Z., L.-H. C. and B.-L. S.; supervision, J. Z., L.-H. C. and B.-L. S.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

L.-H. Chen acknowledges Hubei Provincial Department of Education for the “Chutian Scholar” program. This work is financially supported by National Natural Science Foundation of China (U1663225 and U20A20122), Program for Changjiang Scholars and Innovative Research Team in University (IRT_15R52) of the Chinese Ministry of Education and Program of Introducing Talents of Discipline to Universities-Plan 111 (Grant No. B20002) from the Ministry of Science and Technology and the Ministry of Education of China. This research is also supported by the European Commission Interreg V France-Wallonie-Vlaanderen project “DepollutAir”.

References

  1. F. Akira and H. Kenichi, Electrochemical Photolysis of Water at a Semiconductor Electrode, Nature, 1972, 238, 37–38 CrossRef PubMed.
  2. Q. C. Wei, Y. Chen, Z. Wang, D. Z. Yu, W. H. Wang, J. Q. Li, L.-H. Chen, Y. Li and B.-L. Su, Light–Assisted Semi-Hydrogenation of 1, 3-Butadiene with Water, Angew. Chem., Int. Ed., 2022, 61, e202210573 CAS.
  3. H. Zhao, C. F. Li, L. Y. Liu, B. Palma, Z. Y. Hu, S. Renneckar, S. Larter, Y. Li, M. G. Kibria, J. Hu and B.-L. Su, n-p Heterojunction of TiO2-NiO core-shell structure for efficient hydrogen generation and lignin photoreforming, J. Colloid Interface Sci., 2021, 585, 694–704 CrossRef CAS PubMed.
  4. M. Yang, Y. X. Guo, Z. Liu, X. Y. Li, Q. Huang, X. Y. Yang, C.-F. Ye, Y. Li, J.-P. Liu, L.-H. Chen, B.-L. Su and Y. L. Wang, Engineering Rich Active Sites and Efficient Water Dissociation for Ni-Doped MoS2/CoS2 Hierarchical Structures toward Excellent Alkaline Hydrogen Evolution, Langmuir, 2023, 39, 236–248 CrossRef CAS PubMed.
  5. H. Zhao, J. Liu, C. F. Li, X. Zhang, Y. Li, Z. Y. Hu, B. Li, Z. Chen, J. Hu and B.-L. Su, Meso–microporous nanosheet–constructed 3DOM perovskites for remarkable photocatalytic hydrogen production, Adv. Funct. Mater., 2022, 32, 2112831 CrossRef CAS.
  6. S. Maitra, S. Halder, T. Maitra and S. Roy, Superior light absorbing CdS/vanadium sulphide nanowalls@TiO2 nanorod ternary heterojunction photoanodes for solar water splitting, New J. Chem., 2021, 45, 7353–7367 RSC.
  7. T. W. Wang, Z. W. Yin, Y. H. Guo, F. Y. Bai, J. Chen, W. Dong, J. Liu, Z.-Y. Hu, L. Chen, Y. Li and B.-L. Su, Highly Selective Photocatalytic Conversion of Glucose on Holo-Symmetrically Spherical Three-Dimensionally Ordered Macroporous Heterojunction Photonic Crystal, CCS Chem., 2022, 1–16 Search PubMed.
  8. C. Wang, B. Weng, Y. Liao, B. Liu, M. Keshavarz, Y. Ding, H. Huang, D. Verhaeghe, J. A. Steele, W. Feng, B.-L. Su, J. Hofkens and M. B. Roeffaers, Simultaneous photocatalytic H2 generation and organic synthesis over crystalline-amorphous Pd nanocube decorated Cs3Bi2Br9, Chem. Commun., 2022, 58, 10691–10694 RSC.
  9. Y. Ding, S. Maitra, C. Wang, R. Zheng, T. Barakat, S. Roy, L.-H. Chen and B.-L. Su, Bi-functional Cu-TiO2/CuO photocatalyst for large-scale synergistic treatment of waste sewage containing organics and heavy metal ions, Sci. China Mater., 2023, 66, 179–192 CrossRef CAS.
  10. L. Pei, Z. Ma, J. Zhong, W. Li, X. Wen, J. Guo, P. Liu, Z. Zhang, Q. Mao, J. Zhang, S. Yan and Z. Zou, Oxygen Vacancy–Rich Sr2MgSi2O7:Eu2+, Dy3+ Long Afterglow Phosphor as a Round–the–Clock Catalyst for Selective Reduction of CO2 to CO, Adv. Funct. Mater., 2022, 32, 2208565 CrossRef CAS.
  11. Y. Ding, C. Wang, R. Zheng, S. Maitra, G. Zhang, T. Barakat, S. Roy, B.-L. Su and L.-H. Chen, Three-dimensionally ordered macroporous materials for photo/electrocatalytic sustainable energy conversion, solar cell and energy storage, EnergyChem, 2022, 4, 100081 CrossRef CAS.
  12. S. Maitra, A. Sarkar, T. Maitra, S. Halder, K. Kargupta and S. Roy, Solvothermal phase change induced morphology transformation in CdS/CoFe2O4@Fe2O3 hierarchical nanosphere arrays as ternary heterojunction photoanodes for solar water splitting, New J. Chem., 2021, 45(28), 12721–12737 RSC.
  13. X. Y. Li, S. J. Zhu, Y. L. Wang, T. Lian, X. Y. Yang, C. F. Ye, Y. Li, B.-L. Su and L.-H. Chen, Synergistic regulation of S-vacancy of MoS2-based materials for highly efficient electrocatalytic hydrogen evolution, Front. Chem., 2022, 10, 578 Search PubMed.
  14. T. L. Madanu, S. R. Mouchet, O. Deparis, J. Liu, Y. Li and B.-L. Su, Tuning and transferring slow photons from TiO2 photonic crystals to BiVO4 nanoparticles for unprecedented visible light photocatalysis, J. Colloid Interface Sci., 2022, 634, 290–299 CrossRef PubMed.
  15. C. Wang, B. Weng, M. Keshavarz, M. Q. Yang, H. Huang, Y. Ding, F. Lai, I. Aslam, H. Jin, G. Romolini, B.-L. Su, J. A. Steele, J. Hofkens and M. B. Roeffaers, Photothermal Suzuki Coupling Over a Metal Halide Perovskite/Pd Nanocube Composite Catalyst, ACS Appl. Mater. Interfaces, 2022, 14, 17185–17194 CrossRef CAS PubMed.
  16. W. Wang, H. Zhang, Y. Chen and H. Shi, Efficient degradation of tetracycline via coupling of photocatalysis and photo-fenton processes over a 2D/2D α-Fe2O3/g-C3N4 S-scheme heterojunction catalyst, Acta Phys. – Chim. Sin., 2022, 38, 2104030 Search PubMed.
  17. L. Zhou, Y. Li, Y. Zhang, L. Qiu and Y. Xing, A 0D/2D Bi4V2O11/g-C3N4 S-scheme heterojunction with rapid interfacial charges migration for photocatalytic antibiotic degradation, Acta Phys. – Chim. Sin, 2022, 38, 2112027 Search PubMed.
  18. Z. Bao, M. Xing, Y. Zhou, J. Lv, D. Lei, Y. Zhang, Y. Zhang, J. Cai, J. g Wang, Z. Sun, W. Chen, X. Gan, X. Yang, Q. Han, M. Zhang, J. Dai and Y. Wu, Z-Scheme Flower-Like SnO2/g-C3N4 Composite with Sn2+ Active Center for Enhanced Visible–Light Photocatalytic Activity, Adv. Sustainable Syst., 2021, 5, 2100087 CrossRef CAS.
  19. N. Li, J. Ma, Y. Zhang, L. Zhang and T. Jiao, Recent Developments in Functional Nanocomposite Photocatalysts for Wastewater Treatment: A Review, Adv. Sustainable Syst., 2022, 6, 2200106 CrossRef CAS.
  20. X. Gu, T. Chen, J. Lei, Y. Yang, X. Zheng, S. Zhang, Q. Zhu, X. Fu, S. Meng and S. Chen, Self-assembly synthesis of S-scheme g-C3N4/Bi8(CrO4)O11 for photocatalytic degradation of norfloxacin and bisphenol A, Chin. J. Catal., 2022, 43, 2569–2580 CrossRef CAS.
  21. B. Zhang, X. Hu, E. Liu and J. Fan, Novel S-scheme 2D/2D BiOBr/g-C3N4 heterojunctions with enhanced photocatalytic activity, Chin. J. Catal., 2021, 42, 1519–1529 CrossRef CAS.
  22. F. A. Qaraah, S. A. Mahyoub, A. Hezam, A. Qaraah, Q. A. Drmosh and G. Xiu, Construction of 3D flowers-like O-doped g-C3N4-[N-doped Nb2O5/C] heterostructure with direct S-scheme charge transport and highly improved visible-light-driven photocatalytic efficiency, Chin. J. Catal., 2022, 43, 2637–2651 CrossRef CAS.
  23. S. Das, G. Swain, B. P. Mishra and K. Parida, Tailoring the fusion effect of phase-engineered 1T/2H-MoS2 towards photocatalytic hydrogen evolution, New J. Chem., 2022, 46, 14922–14932 RSC.
  24. L. Biswal, S. Nayak and K. Parida, Rationally designed Ti3C2/N, S-TiO2/g-C3N4 ternary heterostructure with spatial charge separation for enhanced photocatalytic hydrogen evolution, J. Colloid Interface Sci., 2022, 621, 254–266 CrossRef CAS PubMed.
  25. Z. Liang, J. Bai, L. Hao, R. Shen, P. Zhang and X. Li, Photodeposition of NiS Cocatalysts on g-C3N4 with Edge Grafting of 4-(1H-Imidazol-2-yl) Benzoic Acid for Highly Elevated Photocatalytic H2 Evolution, Adv. Sustainable Syst., 2023, 7, 2200143 CrossRef CAS.
  26. X. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Carlsson, K. Domen and M. Antonietti, A metal-free polymeric photocatalyst for hydrogen production from water under visible light, Nat. Mater., 2009, 8, 76–80 CrossRef CAS PubMed.
  27. Y. Ding, S. Maitra, D. A. Esteban, S. Bals, H. Vrielinck, T. Barakat, S. Roy, G. V. Tendeloo, J. Liu, Y. Li, A. Vlad and B.-L. Su, Photochemical production of hydrogen peroxide by digging pro-superoxide radical carbon vacancies in porous carbon nitride, Cell Rep. Phys. Sci., 2022, 3, 100874 CrossRef CAS.
  28. M. Yang, R. Lian, X. Zhang, C. Wang, J. Cheng and X. Wang, Photocatalytic cyclization of nitrogen-centered radicals with carbon nitride through promoting substrate/catalyst interaction, Nat. Commun., 2022, 13, 1–9 Search PubMed.
  29. Y. Ding, S. Maitra, C. Wang, R. Zheng, M. Zhang, T. Barakat, S. Roy, J. Liu, Y. Li, T. Hasan and B.-L. Su, Hydrophilic bi-functional B-doped g-C3N4 hierarchical architecture for excellent photocatalytic H2O2 production and photoelectrochemical water splitting, J. Energy Chem., 2022, 70, 236–247 CrossRef CAS.
  30. T. Yuan, L. Sun, Z. Wu, R. Wang, X. Cai, W. Lin, M. Zheng and X. Wang, Mild and metal-free Birch-type hydrogenation of (hetero) arenes with boron carbonitride in water, Nat. Catal., 2022, 5, 1157–1168 CrossRef CAS.
  31. Y. Ding, S. Maitra, C. Wang, S. Halder, R. Zheng, T. Barakat, S. Roy, L.-H. Chen and B.-L. Su, Vacancy defect engineering in semiconductors for solar light–driven environmental remediation and sustainable energy production, Interdiscip. Mater., 2022, 1, 213–255 CrossRef.
  32. S. Maitra, S. Pal, T. Maitra, S. Halder and S. Roy, Solvothermal Etching-Assisted Phase and Morphology Tailoring in Highly Porous CuFe2O4 Nanoflake Photocathodes for Solar Water Splitting, Energy Fuels, 2021, 35, 14087–14100 CrossRef CAS.
  33. Y. Ding, S. Maitra, S. Halder, C. Wang, R. Zheng, T. Barakat, S. Roy, L.-H. Chen and B.-L. Su, Emerging semiconductors and metal-organic-compounds-related photocatalysts for sustainable hydrogen peroxide production, Matter, 2022, 5, 2119–2167 CrossRef CAS.
  34. J. X. Yang, W. B. Yu, C. F. Li, W. D. Dong, L. Q. Jiang, N. Zhou, Z. Zhuang, J. Liu, Z. Hu, H. Zhao, Y. Li, L. Chen, J. Hu and B.-L. Su, PtO nanodots promoting Ti3C2 MXene in situ converted Ti3C2/TiO2 composites for photocatalytic hydrogen production, Chem. Eng. J., 2021, 420, 129695 CrossRef CAS.
  35. H. Zhao, P. Liu, X. Wu, A. Wang, D. Zheng, S. Wang, Z. Chen, S. Larter, Y. Li, B.-L. Su, M. G. Kibria and J. Hu, Plasmon enhanced glucose photoreforming for arabinose and gas fuel co-production over 3DOM TiO2-Au, Appl. Catal., B, 2021, 291, 120055 CrossRef CAS.
  36. H. S. Mohamed, M. Rabia, X. G. Zhou, X. S. Qin, G. Khabiri, M. Shaban, H. A. Younus, S. Taha, Z.-Y. Hu, J. Liu, Y. Li and B.-L. Su, Phase-junction Ag/TiO2 nanocomposite as photocathode for H2 generation, J. Mater. Sci. Technol., 2021, 83, 179–187 CrossRef CAS.
  37. L. Pei, Y. Yuan, W. Bai, T. Li, H. Zhu, Z. Ma, J. Zhong, S. Yan and Z. Zou, In situ-grown island-shaped hollow graphene on TaON with spatially separated active sites achieving enhanced visible-light CO2 reduction, ACS Catal., 2020, 10, 15083–15091 CrossRef CAS.
  38. J. Hu, Y. Lu, X. L. Liu, C. Janiak, W. Geng, S. M. Wu, X.-F. Zhao, L.-Y. Wang, G. Tian, Y. Zhang, B.-L. Su and X. Y. Yang, Photoinduced terminal fluorine and Ti3+ in TiOF2/TiO2 heterostructure for enhanced charge transfer, CCS Chem., 2020, 2, 1573–1581 CrossRef CAS.
  39. Y. Ding, L. Huang, T. Barakat and B.-L. Su, A Novel 3DOM TiO2 based multifunctional photocatalytic and catalytic platform for energy regeneration and pollutants degradation, Adv. Mater. Interfaces, 2021, 8, 2001879 CrossRef CAS.
  40. K. Das, S. Mansingh, R. Mohanty, D. Sahoo, N. Priyadarshini and K. Parida, 0D–2D Fe2O3/Boron-Doped g-C3N4 S-Scheme Exciton Engineering for Photocatalytic H2O2 Production and Photo-Fenton Recalcitrant Pollutant Detoxification: Kinetics, Influencing Factors, and Mechanism, J. Phys. Chem. C, 2023, 127, 22–40 CrossRef CAS.
  41. K. Das, S. Mansingh, D. Sahoo, R. Mohanty and K. Parida, Engineering an oxygen-vacancy-mediated step-scheme charge carrier dynamic coupling WO3−X/ZnFe2O4 heterojunction for robust photo-Fenton-driven levofloxacin detoxification, New J. Chem., 2022, 46, 5785–5798 RSC.
  42. Y. Q. Yan, Y. Chen, Z. Wang, L. H. Chen, H. L. Tang and B.-L. Su, Electrochemistry-assisted selective butadiene hydrogenation with water, Nat. Commun., 2023, 14, 2106 CrossRef CAS PubMed.
  43. H. Zhao, J. Liu, N. Zhong, S. Larter, Y. Li, M. G. Kibria, B.-L. Su, Z. Chen and J. Hu, Biomass Photoreforming for Hydrogen and Value-Added Chemicals Co-Production on Hierarchically Porous Photocatalysts, Adv. Energy Mater., 2023, 2300257 Search PubMed.
  44. B. Mishra and K. Parida, Orienting Z scheme charge transfer in graphitic carbon nitride-based systems for photocatalytic energy and environmental applications, J. Mater. Chem. A, 2021, 9, 10039–10080 RSC.
  45. L. Biswal, L. Acharya, B. Mishra, S. Das, G. Swain and K. Parida, Interfacial Solid-State Mediator-Based Z-Scheme Heterojunction TiO2@Ti3C2/MgIn2S4 Microflower for Efficient Photocatalytic Pharmaceutical Micropollutant Degradation and Hydrogen Generation: Stability, Kinetics, and Mechanistic Insights, ACS Appl. Energy Mater., 2023, 6, 2081–2096 CrossRef CAS.
  46. W. Yu, J. Liu, M. Yi, J. Yang, W. Dong, C. Wang, H. Zhao, H. Mohamed, Z. Wang, L. Chen, Y. Li and B-L. Su, Active faceted Cu2O hollow nanospheres for unprecedented adsorption and visible-light degradation of pollutants, J. Colloid Interface Sci., 2020, 565, 207–217 CrossRef CAS PubMed.
  47. L. Acharya, S. Nayak, S. P. Pattnaik, R. Acharya and K. Parida, Resurrection of boron nitride in p-n type-II boron nitride/B-doped-g-C3N4 nanocomposite during solid-state Z-scheme charge transfer path for the degradation of tetracycline hydrochloride, J. Colloid Interface Sci., 2020, 566, 211–223 CrossRef CAS PubMed.
  48. L. Acharya, S. P. Pattnaik, A. Behera, R. Acharya and K. Parida, Exfoliated boron nitride (e-BN) tailored exfoliated graphitic carbon nitride (e-CN): an improved visible light mediated photocatalytic approach towards TCH degradation and H2 evolution, Inorg. Chem., 2021, 60, 5021–5033 CrossRef CAS PubMed.
  49. J. Yang, Y. Peng, S. Li, J. Mu, Z. Huang, J. Ma, Z. Shi and Q. Jia, Metal nanocluster-based hybrid nanomaterials: Fabrication and application, Coord. Chem. Rev., 2022, 456, 214391 CrossRef CAS.
  50. Y. Wang, Y. X. Chen, T. Barakat, T. M. Wang, A. Krief, Y. J. Zeng, M. Laboureur, L. Fusaro, H. Liao and B.-L. Su, Synergistic effects of carbon doping and coating of TiO2 with exceptional photocurrent enhancement for high performance H2 production from water splitting, J. Energy Chem., 2021, 56, 141–151 CrossRef CAS.
  51. S. Das, G. Swain and K. Parida, One step towards the 1T/2H-MoS2 mixed phase: a journey from synthesis to application, Mater. Chem. Front., 2021, 5, 2143–2172 RSC.
  52. Y. Lu, Y. X. Liu, L. He, L. Y. Wang, X. L. Liu, J. W. Liu and B.-L. Su, Interfacial co-existence of oxygen and titanium vacancies in nanostructured TiO2 for enhancement of carrier transport, Nanoscale, 2020, 12, 8364–8370 RSC.
  53. A. Balakrishnan and M. Chinthala, Comprehensive review on advanced reusability of g-C3N4 based photocatalysts for the removal of organic pollutants, Chemosphere, 2022, 297, 134190 CrossRef CAS PubMed.
  54. L. Biswal, S. Nayak and K. Parida, Recent progress on strategies for the preparation of 2D/2D MXene/g-C3N4 nanocomposites for photocatalytic energy and environmental applications, Catal. Sci. Technol., 2021, 11, 1222–1248 RSC.
  55. L. Acharya, B. P. Mishra, S. P. Pattnaik, R. Acharya and K. Parida, Incorporating nitrogen vacancies in exfoliated B-doped g-C3N4 towards improved photocatalytic ciprofloxacin degradation and hydrogen evolution, New J. Chem., 2022, 46, 3493–3503 RSC.
  56. X. Liu, R. Ma, L. Zhuang, B. Hu, J. Chen, X. Liu and X. Wang, Recent developments of doped g-C3N4 photocatalysts for the degradation of organic pollutants, Crit. Rev. Environ. Sci. Technol., 2021, 51, 751–790 CrossRef CAS.
  57. H. Chen, Y. Xie, X. Sun, M. Lv, F. Wu, L. Zhang, L. Li and X. Xu, Efficient charge separation based on type-II g-C3N4/TiO2-B nanowire/tube heterostructure photocatalysts, Dalton Trans., 2015, 44, 13030–13039 RSC.
  58. X. Han, D. Xu, L. An, C. Hou, Y. Li, Q. Zhang and H. Wang, WO3/g-C3N4 two-dimensional composites for visible-light driven photocatalytic hydrogen production, Int. J. Hydrogen Energy, 2018, 43, 4845–4855 CrossRef CAS.
  59. A. Bhosale, A. Gophane, J. Kadam, S. Sabale, K. Sonawane and K. Garadkar, Fabrication of visible-active ZnO-gC3N4 nanocomposites for photodegradation and cytotoxicity of methyl orange and antibacterial activity towards drug resistance pathogens, Opt. Mater., 2023, 136, 113392 CrossRef CAS.
  60. X. Zheng, G. Shen, C. Wang, Y. Li, D. Dunphy, T. Hasan, C. J. Brinker and B.-L. Su, Bio-inspired Murray materials for mass transfer and activity, Nat. Commun., 2017, 8, 1–9 CrossRef.
  61. V. Soni, P. Singh, A. A. P. Khan, A. Singh, A. K. Nadda, C. M. Hussain and P. Raizada, Photocatalytic transition-metal-oxides-based p-n heterojunction materials: synthesis, sustainable energy and environmental applications, and perspectives, J. Nanostruct. Chem., 2023, 13, 129–166 CrossRef CAS.
  62. H. Zhao, M. Zalfani, C. F. Li, J. Liu, Z. Y. Hu, M. Mahdouani, R. Bourguig, Y. Li and B.-L. Su, Cascade electronic band structured zinc oxide/bismuth vanadate/three-dimensional ordered macroporous titanium dioxide ternary nanocomposites for enhanced visible light photocatalysis, J. Colloid Interface Sci., 2019, 539, 585–597 CrossRef CAS PubMed.
  63. A. Karmakar, E. Velasco and J. Li, Metal-organic frameworks as effective sensors and scavengers for toxic environmental pollutants, Natl. Sci. Rev., 2022, 9, nwac091 CrossRef CAS PubMed.
  64. Y. Yin, X. Kang and B. Han, Two-dimensional materials: synthesis and applications in the electro-reduction of carbon dioxide, Chem. Synth., 2022, 2, 19 CrossRef.
  65. M.-H. Sun, S.-Z. Huang, L.-H. Chen, Y. Li, X.-Y. Yang, Z.-Y. Yuan and B.-L. Su, Applications of hierarchically structured porous materials from energy storage and conversion, catalysis, photocatalysis, adsorption, separation, and sensing to biomedicine, Chem. Soc. Rev., 2016, 45, 3479–3563 RSC.
  66. J. Yu, W. Wang, B. Cheng and B. L. Su, Enhancement of photocatalytic activity of mesporous TiO2 powders by hydrothermal surface fluorination treatment, J. Phys. Chem. C, 2009, 113, 6743–6750 CrossRef CAS.
  67. H. Wang, L. Wang, Q. Luo, J. Zhang, C. Wang, X. Ge and F.-S. Xiao, Two-dimensional manganese oxide on ceria for the catalytic partial oxidation of hydrocarbons, Chem. Synth., 2022, 2, 2 CrossRef.
  68. Y. Li, D. Zhang, W. Qiao, H. Xiang, F. Besenbacher, Y. Li and R. Su, Nanostructured heterogeneous photocatalyst materials for green synthesis of valuable chemicals, Chem. Synth., 2022, 2, 9 CrossRef.
  69. L. H. Chen, X. Y. Li, J. C. Rooke, Y. H. Zhang, X. Y. Yang, Y. Tang, F.-S. Xiao and B.-L. Su, Hierarchically structured zeolites: synthesis, mass transport properties and applications, J. Mater. Chem., 2012, 22, 17381–17403 RSC.
  70. L. Wu, Y. Li, Z. Fu and B.-L. Su, Hierarchically structured porous materials: Synthesis strategies and applications in energy storage, Natl. Sci. Rev., 2020, 7, 1667–1701 CrossRef CAS PubMed.
  71. A. Ajmal, I. Majeed, R. N. Malik, H. Idriss and M. A. Nadeem, Principles and mechanisms of photocatalytic dye degradation on TiO2 based photocatalysts: a comparative overview, RSC Adv., 2014, 4, 37003–37026 RSC.
  72. M. Zalfani, B. Schueren, M. Mahdouani, R. Bourguiga, W. B. Yu, M. Wu, O. Deparis, Y. Li and B.-L. Su, ZnO quantum dots decorated 3DOM TiO2 nanocomposites: Symbiose of quantum size effects and photonic structure for highly enhanced photocatalytic degradation of organic pollutants, Appl. Catal., B, 2016, 199, 187–198 CrossRef CAS.
  73. M. Rochkind, S. Pasternak and Y. Paz, Using dyes for evaluating photocatalytic properties: a critical review, Molecules, 2014, 20, 88–110 CrossRef PubMed.
  74. S. Sakthivel, B. Neppolian, M. V. Shankar, B. Arabindoo, M. Palanichamy and V. Murugesan, Solar photocatalytic degradation of azo dye: comparison of photocatalytic efficiency of ZnO and TiO2, Sol. Energy Mater. Sol. Cells, 2003, 77, 65–82 CrossRef CAS.
  75. M. Zalfani, B. V. Schueren, Z.-Y. Hu, J. C. Rooke, R. Bourguiga, M. Wu, Y. Li, G. V. Tendeloo and B.-L. Su, Novel 3DOM BiVO4/TiO2 nanocomposites for highly enhanced photocatalytic activity, J. Mater. Chem. A, 2015, 3, 21244–21256 RSC.
  76. Q. Li, N. Zhang, Y. Yang, G. Wang and D. H. Ng, High efficiency photocatalysis for pollutant degradation with MoS2/C3N4 heterostructures, Langmuir, 2014, 30, 8965–8972 CrossRef CAS PubMed.
  77. Y. Sheng, Z. Wei, H. Miao, W. Yao, H. Li and Y. Zhu, Enhanced organic pollutant photodegradation via adsorption/photocatalysis synergy using a 3D g-C3N4/TiO2 free-separation photocatalyst, Chem. Eng. J., 2019, 370, 287–294 CrossRef CAS.
  78. R. Hao, G. Wang, H. Tang, L. Sun, C. Xu and D. Han, Template-free preparation of macro/mesoporous g-C3N4/TiO2 heterojunction photocatalysts with enhanced visible light photocatalytic activity, Appl. Catal., B, 2016, 187, 47–58 CrossRef CAS.
  79. P. Xia, B. Zhu, B. Cheng, J. Yu and J. Xu, 2D/2D g-C3N4/MnO2 nanocomposite as a direct Z-scheme photocatalyst for enhanced photocatalytic activity, ACS Sustainable Chem. Eng., 2018, 6, 965–973 CrossRef CAS.
  80. Y. Ren, W. Zhan, L. Tang, H. Zheng, H. Liu and K. Tang, Constructing a ternary H2SrTa2O7/gC3N4/Ag3PO4 heterojunction based on cascade electron transfer with enhanced visible light photocatalytic activity, CrystEngComm, 2020, 22, 6485–6494 RSC.
  81. S. J. Liu, F. T. Li, Y. L. Li, Y. J. Hao, X. J. Wang, B. Li and R. H. Liu, Fabrication of ternary g-C3N4/Al2O3/ZnO heterojunctions based on cascade electron transfer toward molecular oxygen activation, Appl. Catal., B, 2017, 212, 115–128 CrossRef CAS.
  82. J. Xiong, X. Li, J. Huang, X. Gao, Z. Chen, J. Liu and Y. Zhu, CN/rGO@ BPQDs high-low junctions with stretching spatial charge separation ability for photocatalytic degradation and H2O2 production, Appl. Catal., B, 2020, 266, 118602 CrossRef CAS.
  83. R. Khawaja, S. Sonar, T. Barakat, N. Heymans, B.-L. Su, A. Löfberg and S. Siffert, VOCs catalytic removal over hierarchical porous zeolite NaY supporting Pt or Pd nanoparticles, Catal. Today, 2022, 405, 212–220 CrossRef.
  84. T. Barakat, J. C. Rooke, R. Cousin, J. F. Lamonier, J. M. Giraudon, B.-L. Su and S. Siffert, Investigation of the elimination of VOC mixtures over a Pd-loaded V-doped TiO2 support, New J. Chem., 2014, 38, 2066–2074 RSC.
  85. E. Haye, T. Barakat, L. Chavee, B.-L. Su, S. Lucas, L. Houssiau and A. Achour, Low-pressure plasma process for the dry synthesis of cactus-like Au-TiO2 nanocatalysts for toluene degradation, App. Surf. Sci., 2022, 571, 151313 CrossRef CAS.
  86. J. C. Rooke, T. Barakat, J. Brunet, Y. Li, M. F. Finol, J. F. Lamonier and B.-L. Su, Hierarchically nanostructured porous group Vb metal oxides from alkoxide precursors and their role in the catalytic remediation of VOCs, Appl. Catal., B, 2015, 162, 300–309 CrossRef CAS.
  87. T. Barakat, V. Idakiev, R. Cousin, G. S. Shao, Z. Y. Yuan, T. Tabakova and S. Siffert, Total oxidation of toluene over noble metal based Ce, Fe and Ni doped titanium oxides, Appl. Catal., B, 2014, 146, 138–146 CrossRef CAS.
  88. J. C. Rooke, T. Barakat, M. F. Finol, P. Billemont, G. D. Weireld, Y. Li and B.-L. Su, Influence of hierarchically porous niobium doped TiO2 supports in the total catalytic oxidation of model VOCs over noble metal nanoparticles, Appl. Catal., B, 2013, 142, 149–160 CrossRef.
  89. Y. Wang, W. Jiang, W. Luo, X. Chen and Y. Zhu, Ultrathin nanosheets g-C3N4@ Bi2WO6 core-shell structure via low temperature reassembled strategy to promote photocatalytic activity, Appl. Catal., B, 2018, 237, 633–640 CrossRef CAS.
  90. S. Obregón, Y. Zhang and G. Colón, Cascade charge separation mechanism by ternary heterostructured BiPO4/TiO2/g-C3N4 photocatalyst, Appl. Catal., B, 2016, 184, 96–103 CrossRef.
  91. S. S. Kerur, S. Bandekar, M. S. Hanagadakar, S. S. Nandi, G. M. Ratnamala and P. G. Hegde, Removal of hexavalent Chromium-Industry treated water and Wastewater: A review, Mater. Today: Proc., 2021, 42, 1112–1121 CAS.
  92. S. Loyaux-Lawniczak, P. Lecomte and J. J. Ehrhardt, Behavior of hexavalent chromium in a polluted groundwater: redox processes and immobilization in soils, Environ. Sci. Technol., 2001, 35, 1350–1357 CrossRef CAS PubMed.
  93. X. Deng, Y. Chen, J. Wen, Y. Xu, J. Zhu and Z. Bian, Polyaniline-TiO2 composite photocatalysts for light-driven hexavalent chromium ions reduction, Sci. Bull., 2020, 65, 105–112 CrossRef CAS PubMed.
  94. G. Ghosh and P. K. Bhattacharya, Hexavalent chromium ion removal through micellar enhanced ultrafiltration, Chem. Eng. J., 2006, 119, 45–53 CrossRef CAS.
  95. N. A. Azeez, S. S. Dash, S. N. Gummadi and V. S. Deepa, Nano-remediation of toxic heavy metal contamination: Hexavalent chromium [Cr(VI)], Chemosphere, 2021, 266, 129204 CrossRef CAS PubMed.
  96. Q. Zhong, H. Lan, M. Zhang, H. Zhu and M. Bu, Preparation of heterostructure g-C3N4/ZnO nanorods for high photocatalytic activity on different pollutants (MB, RhB, Cr(VI) and eosin), Ceram. Interfaces, 2020, 46, 12192–12199 CrossRef CAS.
  97. C. V. Reddy, R. Koutavarapu, J. Shim, B. Cheolho and K. R. Reddy, Novel g-C3N4/Cu-doped ZrO2 hybrid heterostructures for efficient photocatalytic Cr(VI) photoreduction and electrochemical energy storage applications, Chemosphere, 2022, 295, 133851 CrossRef CAS.
  98. Y. Wang, S. Bao, Y. Liu, W. Yang, Y. Yu, M. Feng and K. Li, Efficient photocatalytic reduction of Cr(VI) in aqueous solution over CoS2/g-C3N4-rGO nanocomposites under visible light, App. Surf. Sci., 2020, 510, 145495 CrossRef CAS.
  99. P. Babu, S. Mohanty, B. Naik and K. Parida, Serendipitous assembly of mixed phase BiVO4 on B-doped g-C3N4: an appropriate p-n heterojunction for photocatalytic O2 evolution and Cr(VI) reduction, Inorg. Chem., 2019, 58, 12480–12491 CrossRef CAS PubMed.
  100. S. Nayak and K. Parida, Dynamics of charge-transfer behavior in a plasmon-induced quasi-type-II p–n/n–n dual heterojunction in Ag@Ag3PO4/g-C3N4/NiFe LDH nanocomposites for photocatalytic Cr(VI) reduction and phenol oxidation, ACS Omega, 2018, 3, 7324–7343 CrossRef CAS PubMed.
  101. X. H. Yi, S. Q. Ma, X. D. Du, C. Zhao, H. Fu, P. Wang and C. C. Wang, The facile fabrication of 2D/3D Z-scheme g-C3N4/UiO-66 heterojunction with enhanced photocatalytic Cr(VI) reduction performance under white light, Chem. Eng. J., 2019, 375, 121944 CrossRef CAS.
  102. K. Gillingham and J. H. Stock, The cost of reducing greenhouse gas emissions, J. Econ. Perspect., 2018, 32, 53–72 CrossRef.
  103. D. Liu, X. Guo and B. Xiao, What causes growth of global greenhouse gas emissions? Evidence from 40 countries, Sci. Total Environ., 2019, 661, 750–766 CrossRef CAS PubMed.
  104. I. Arto and E. Dietzenbacher, Drivers of the growth in global greenhouse gas emissions, Environ. Sci. Technol., 2014, 48, 5388–5394 CrossRef CAS PubMed.
  105. N. N. Vu, S. Kaliaguine and T. O. Do, Plasmonic Photocatalysts for Sunlight–Driven Reduction of CO2: Details, Developments, and Perspectives, ChemSusChem, 2020, 13, 3967–3991 CrossRef CAS PubMed.
  106. N. N. Vu, S. Kaliaguine and T. O. Do, Critical aspects and recent advances in structural engineering of photocatalysts for sunlight–driven photocatalytic reduction of CO2 into fuels, Adv. Funct. Mater., 2019, 29, 1901825 CrossRef.
  107. D. Grammatico, H. N. Tran, Y. Li, S. Pugliese, L. Billon, B.-L. Su and M. Fontecave, Immobilization of a Molecular Re Complex on MOF–derived Hierarchical Porous Carbon for CO2 Electroreduction in Water/Ionic Liquid Electrolyte, ChemSusChem, 2020, 13, 6418–6425 CAS.
  108. E. B. Stechel and J. E. Miller, Re-energizing CO2 to fuels with the sun: Issues of efficiency, scale, and economics, J. CO2 Util., 2013, 1, 28–36 CrossRef CAS.
  109. D. Grammatico, A. J. Bagnall, L. Riccardi, M. Fontecave, B.-L. Su and L. Billon, Heterogenised molecular catalysts for sustainable electrochemical CO2 reduction, Angew. Chem., Int. Ed., 2022, 61, e202206399 CrossRef CAS PubMed.
  110. S. Pugliese, N. T. Huan, J. Forte, D. Grammatico, S. Zanna, B.-L. Su, Y. Li and M. Fontecave, Functionalization of Carbon Nanotubes with Nickel Cyclam for the Electrochemical Reduction of CO2, ChemSusChem, 2020, 13, 6449–6456 CAS.
  111. X. Li, J. Yu, M. Jaroniec and X. Chen, Cocatalysts for selective photoreduction of CO2 into solar fuels, Chem. Rev., 2019, 119, 3962–4179 CrossRef CAS PubMed.
  112. C. Aprile, F. Giacalone, P. Agrigento, L. F. Liotta, J. A. Martens, P. P. Pescarmona and M. Gruttadauria, Multilayered Supported Ionic Liquids as Catalysts for Chemical Fixation of Carbon Dioxide: A High–Throughput Study in Supercritical Conditions, ChemSusChem, 2011, 4, 1830–1837 CrossRef CAS PubMed.
  113. R. Li, W. Zhang and K. Zhou, Metal-organic–framework–based catalysts for photoreduction of CO2, Adv. Mater., 2018, 30, 1705512 CrossRef PubMed.
  114. C. Calabrese, F. Giacalone and C. Aprile, Hybrid catalysts for CO2 conversion into cyclic carbonates, Catalysts, 2019, 9, 325 CrossRef.
  115. P. Agrigento, S. M. Al-Amsyar, B. Sorée, M. Taherimehr, M. Gruttadauria, C. Aprile and P. P. Pescarmona, Synthesis and high-throughput testing of multilayered supported ionic liquid catalysts for the conversion of CO2 and epoxides into cyclic carbonates, Catal. Sci. Technol., 2014, 4, 1598–1607 RSC.
  116. M. Zhou, S. Wang, P. Yang, C. Huang and X. Wang, Boron carbon nitride semiconductors decorated with CdS nanoparticles for photocatalytic reduction of CO2, ACS Catal., 2018, 8, 4928–4936 CrossRef CAS.
  117. S. W. Cao, X. F. Liu, Y. P. Yuan, Z. Y. Zhang, Y. S. Liao, J. Fang and C. Xue, Solar-to-fuels conversion over In2O3/g-C3N4 hybrid photocatalysts, Appl. Catal., B, 2014, 147, 940–946 CrossRef CAS.
  118. H. Shi, G. Chen, C. Zhang and Z. Zou, Polymeric g-C3N4 coupled with NaNbO3 nanowires toward enhanced photocatalytic reduction of CO2 into renewable fuel, ACS Catal., 2014, 4, 3637–3643 CrossRef CAS.
  119. R. Bhosale, S. Jain, C. P. Vinod, S. Kumar and S. Ogale, Direct Z-scheme g-C3N4/FeWO4 nanocomposite for enhanced and selective photocatalytic CO2 reduction under visible light, ACS Appl. Mater. Interfaces, 2019, 11, 6174–6183 CrossRef CAS PubMed.
  120. T. Di, B. Zhu, B. Cheng, J. Yu and J. Xu, A direct Z-scheme g-C3N4/SnS2 photocatalyst with superior visible-light CO2 reduction performance, J. Catal., 2017, 352, 532–541 CrossRef CAS.
  121. N. Nie, L. Zhang, J. Fu, B. Cheng and J. Yu, Self-assembled hierarchical direct Z-scheme g-C3N4/ZnO microspheres with enhanced photocatalytic CO2 reduction performance, App. Surf. Sci., 2018, 441, 12–22 CrossRef CAS.
  122. Y. He, L. Zhang, B. Teng and M. Fan, New application of Z-scheme Ag3PO4/g-C3N4 composite in converting CO2 to fuel, Environ. Sci. Technol., 2015, 49, 649–656 CrossRef CAS PubMed.
  123. L. Cheng, Q. Xiang, Y. Liao and H. Zhang, CdS-based photocatalysts, Energy Environ. Sci., 2018, 11, 1362–1391 RSC.
  124. M. A. Correa-Duarte, M. Giersig and L. M. Liz-Marzán, Stabilization of CdS semiconductor nanoparticles against photodegradation by a silica coating procedure, Chem. Phys. Lett., 1998, 286, 497–501 CrossRef CAS.
  125. Q. Li, X. Li, S. Wageh, A.-A. Al-Ghamdi and J. Yu, CdS/graphene nanocomposite photocatalysts, Adv. Energy Mater., 2015, 5, 1500010 CrossRef.
  126. Y. Ding, C. Wang, J. Zhong, Q. Mao, R. Zheng, Y. H. Ng, M.-H. Sun, S. Maitra, L.-H. Chen and B.-L. Su, Three-dimensionally ordered macroporous materials for pollutants abatement, environmental sensing and bacterial inactivation, Sci. China: Chem., 2023 DOI:10.1007/s11426-023-1574-1.
  127. L. Wang, B.-B. Zhang, X.-Y. Yang and B.-L. Su, Alginate@ polydopamine@SiO2 microcapsules with controlled porosity for whole-cell based enantioselective biosynthesis of (S)-1-phenylethanol, Colloids Surf., B, 2022, 214, 112454 CrossRef CAS PubMed.
  128. L. Wang, Y. Li, X. Y. Yang, B. B. Zhang, N. Ninane, H. J. Busscher, Z. Hu, C. Delneuville, N. Jiang, H. Xie, G. Tendeloo, T. Hasan and B.-L. Su, Single-cell yolk-shell nanoencapsulation for long-term viability with size-dependent permeability and molecular recognition, Natl. Sci. Rev., 2021, 8, nwaa097 CrossRef CAS PubMed.
  129. P. Xu, M. L. Janex, P. Savoye, A. Cockx and V. Lazarova, Wastewater disinfection by ozone: main parameters for process design, Water Res., 2022, 36, 1043–1055 CrossRef PubMed.
  130. A. Zhang, H. Xie, N. Liu, B.-L. Chen, H. Ping, Z.-Y. Fu and B.-L. Su, Crystallization of calcium carbonate under the influences of casein and magnesium ions, RSC Adv., 2016, 6, 110362–110366 RSC.
  131. A. Simpson and W. Mitch, Chlorine and ozone disinfection and disinfection byproducts in postharvest food processing facilities: A review, Crit. Rev. Environ. Sci. Technol., 2022, 52, 1825–1867 CrossRef.
  132. C. Regmi, B. Joshi, S. K. Ray, G. Gyawali and R. P. Pandey, Understanding mechanism of photocatalytic microbial decontamination of environmental wastewater, Front. Chem., 2018, 6, 33 CrossRef PubMed.
  133. O. K. Dalrymple, E. Stefanakos, M. A. Trotz and D. Y. Goswami, A review of the mechanisms and modeling of photocatalytic disinfection, Appl. Catal., B, 2010, 98, 27–38 CrossRef CAS.
  134. S. H. Xue, H. Xie, H. Ping, Q. C. Li, B.-L. Su and Z.-Y. Fu, Induced transformation of amorphous silica to cristobalite on bacterial surfaces, RSC Adv., 2015, 5, 71844–71848 RSC.
  135. W. Wang, G. Huang, C. Y. Jimmy and P. K. Wong, Advances in photocatalytic disinfection of bacteria: development of photocatalysts and mechanisms, J. Environ. Sci., 2015, 34, 232–247 CrossRef CAS PubMed.
  136. J. Deng, J. Liang, M. Li and M. Tong, Enhanced visible-light-driven photocatalytic bacteria disinfection by g-C3N4-AgBr, Colloids Surf., B, 2017, 152, 49–57 CrossRef CAS PubMed.
  137. X. Geng, L. Wang, L. Zhang, H. Wang, Y. Peng and Z. Bian, H2O2 production and in situ sterilization over a ZnO/g-C3N4 heterojunction photocatalyst, Chem. Eng. J., 2021, 420, 129722 CrossRef CAS.
  138. H. Gourama and L. B. Bullerman, Aspergillus flavus and Aspergillus parasiticus: Aflatoxigenic fungi of concern in foods and feeds: A review, J. Food Prot., 1995, 58, 1395–1404 CrossRef CAS PubMed.
  139. B. B. Zhang, L. Wang, V. Charles, J. C. Rooke and B.-L. Su, Robust and biocompatible hybrid matrix with controllable permeability for microalgae encapsulation, ACS Appl. Mater. Interfaces, 2016, 8, 8939–8946 CrossRef CAS PubMed.
  140. D. Sun, J. Mao, L. Cheng, X. Yang, H. Li, L. Zhang, W. Zhang, Q. Zhang and P. Li, Magnetic g-C3N4/NiFe2O4 composite with enhanced activity on photocatalytic disinfection of Aspergillus flavus, Chem. Eng. J., 2021, 418, 129417 CrossRef CAS.

This journal is © the Partner Organisations 2023