Next Article in Journal
Polypyrrole-Coated Low-Crystallinity Iron Oxide Grown on Carbon Cloth Enabling Enhanced Electrochemical Supercapacitor Performance
Next Article in Special Issue
Synthesis and Application of Liquid Metal Based-2D Nanomaterials: A Perspective View for Sustainable Energy
Previous Article in Journal
Advanced Fitting Method for the Kinetic Analysis of Thermogravimetric Data
Previous Article in Special Issue
Characterization of Fe3O4 Nanoparticles for Applications in Catalytic Activity in the Adsorption/Degradation of Methylene Blue and Esterification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in g-C3N4-Based Materials and Their Application in Energy and Environmental Sustainability

1
School of Chemistry and Chemical Engineering, Xi’an Shiyou University, Xi’an 710065, China
2
Xi’an Key Laboratory of Low-Carbon Utilization for High-Carbon Resources, Xi’an Shiyou University, Xi’an 710065, China
3
State Key Laboratory of Petroleum Pollution Control, Xi’an Shiyou University, Xi’an 710065, China
4
National Center for Quality Supervision and Inspection of Oil and Gas Products (Yan’an), Yan’an 716000, China
5
Department of Crop Soil Sciences, Washington State University, Pullman, WA 99164-6420, USA
6
Department of Statistics, North Dakota State University, Fargo, ND 58105, USA
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(1), 432; https://doi.org/10.3390/molecules28010432
Submission received: 5 December 2022 / Revised: 22 December 2022 / Accepted: 25 December 2022 / Published: 3 January 2023

Abstract

:
Graphitic carbon nitride (g-C3N4), with facile synthesis, unique structure, high stability, and low cost, has been the hotspot in the field of photocatalysis. However, the photocatalytic performance of g-C3N4 is still unsatisfactory due to insufficient capture of visible light, low surface area, poor electronic conductivity, and fast recombination of photogenerated electron-hole pairs. Thus, different modification strategies have been developed to improve its performance. In this review, the properties and preparation methods of g-C3N4 are systematically introduced, and various modification approaches, including morphology control, elemental doping, heterojunction construction, and modification with nanomaterials, are discussed. Moreover, photocatalytic applications in energy and environmental sustainability are summarized, such as hydrogen generation, CO2 reduction, and degradation of contaminants in recent years. Finally, concluding remarks and perspectives on the challenges, and suggestions for exploiting g-C3N4-based photocatalysts are presented. This review will deepen the understanding of the state of the art of g-C3N4, including the fabrication, modification, and application in energy and environmental sustainability.

1. Introduction

Along with the rapid growth of the global population and the development of industrialization and urbanization, the demand for fossil energy, such as petroleum, coal, and natural gas, is increasing, as well as the deterioration of environmental pollution [1]. Driven by the ongoing energy and ecological crisis, the development of sustainable energy is a matter of great urgency and is related to the vital interests of people worldwide. As a green and renewable energy, solar energy has become a hot topic [2] and converting it into chemical energy by using photocatalysis is regarded as a potential pathway to supply renewable energy and alleviate environmental issues in the future. The photocatalyst is the most important key in economic photocatalysis application; it should be efficient, stable, low-cost, and capable of harvesting visible light [3]. Many photocatalytic materials have been reported and used in various fields, including hydrogen evolution, contaminant photo-oxidization or photodecomposition, and photoelectrochemical conversion [4,5]. For example, titanium dioxide (TiO2) and related photocatalysts have been used in solar energy conversion, due to their merits of low price, unique optical-electronic properties, great durability, and non-toxicity [6]. However, the large bandgap of TiO2 (3.2 eV) prevents it from actual solar energy utilization [7,8]. Compared to TiO2, the newly emerged graphitic carbon nitride (g-C3N4) is a visible-light-response material with a narrower bandgap of 2.7 eV. It is a kind of metal-free photocatalyst and possesses the advantages of simple synthesis, suitable semi-conducting properties, and high structural stability under both thermal and photochemical conditions. These merits make g-C3N4 a unique material for energy and environmental applications, including photocatalytic H2 generation, CO2 reduction, and degradation of organic pollutants (dyes, pesticides, pharmaceuticals, phenolic compounds) and inorganic pollutants like heavy metals, carbon dioxide reduction, etc. [6,9,10,11,12,13,14]. In 2009, Wang et al. [15] for the first time proved that g-C3N4 could be used for photocatalytic hydrogen production upon visible-light irradiation, marking a significant milestone in metal-free photocatalysts. Nevertheless, due to the rapid recombination of photo-induced charges, the photocatalytic performance of g-C3N4 still possesses a significant possibility to be further enhanced, and the research on g-C3N4 gives rise to a new upsurge. So far, different strategies have been adopted to increase the photocatalytic efficiency of g-C3N4, such as morphological control, element doping, heterojunction construction, and nanomaterial composition [8,16,17]. In this review, we introduce the recent advances in the preparation and modification of g-C3N4, and summarize its photocatalytic application in H2 generation, CO2 reduction, and degradation of organic pollutants in recent years. Finally, we conclude the research challenges with g-C3N4 and suggest perspectives for future research direction.

2. Properties and Preparation of g-C3N4

2.1. The Origin and Properties of g-C3N4

Carbon nitride (C3N4), composed of carbon and nitrogen elements, is a kind of organic semiconductor material. The history of C3N4 can be traced back to 1834. One type of material named “melon” was synthesized by Berzelius and also reported by Liebig, which is a linear polymer connected triazine and tri-s-triazines (Figure 1) via secondary nitrogen [18,19]. However, this material did not attract much attention due to the lack of comprehensive characterization at the time. Along with the development of characterization methods, Franklin [20] probed this material in 1922 and proposed the concept of C3N4, which indicated that C3N4 could be obtained by polymerizing various ammonia carbonic acids. Later, Pauling and Sturdivant [21] demonstrated that tri-s-triazine was the unit of melon in 1937. Still, its chemical instability and insolubility in most reagents made it impossible to unveil the structure until 1989, researchers found that when Si in β-Si3N4 was replaced by C, the derived β-C3N4 was as hard as a diamond [22]. Based on the study, five types of C3N4, comprising α, β, pseudocubic, cubic, and graphitic, were predicated in 1996 [23]. Among them, the first four materials are hard materials but not favorable to be synthesized due to their low stability [24] and g-C3N4 is confirmed as the most resistant under surrounding situations; it possesses a similar layered structure to graphene and sp2 hybrid π-conjugated electronic band structure. In recent years, large numbers of g-C3N4 materials have been synthesized by thermal polymerization of urea, melamine, cyanamide, dicyandiamide, and thiourea, which indicated that g-C3N4 was composed of melem units and further confirmed that tri-s-triazine was the basic unit of g-C3N4.
The electronic structures of both carbon and nitrogen atoms in g-C3N4 determine their electronic and optical properties. The lone pair electrons of carbon and nitrogen atoms of g-C3N4 interact to create a large π bond, analogous to the benzene ring, and then form a highly delocalized conjugate system. The delocalized conjugate chemical structure contributes to the formation of the stacking of the carbon nitride layer, which connects through amines, and such a structure makes the superior electronic conductivity of g-C3N4 [1]. In addition, the solid covalent bond between carbon and nitrogen atoms leads to the excellent chemical and thermal stability of g-C3N4. After experimental measurements [26,27] the conduction band (CB) and valence band (VB) of g-C3N4 are −1.3 V and 1.4 V at pH = 7 versus the standard hydrogen electrode (SHE), respectively. Such band positions promote visible light harvesting under oxidation and reduction systems. In short, g-C3N4 possesses unique electronic, structural, physicochemical, and optical properties, sufficient for photocatalytic application in H2 production, CO2 photoreduction, and degradation of organic pollutants [9,28].

2.2. Preparation of g-C3N4

The property of a catalyst varies substantially depending upon the preparation protocols. To achieve the application of g-C3N4 in the field of photocatalysis, the synthesis of high-performance g-C3N4 is a prerequisite. Various methods have been proposed, including thermal condensation, hydrothermal and solvothermal approaches, solid-state fabrication, self-hand synthesis, template-supported formation, deposition-precipitation, and ball milling process [1,29].
Among these methods, thermal condensation, a combination of polycondensation and polyaddition, is the most common method to prepare g-C3N4. Nitrogen-rich chemicals, such as melamine, urea, and dicyandiamide, are usually used as precursors, and after the deamination process, g-C3N4 is generated under high temperatures. For example, Yan et al. [30] synthesized g-C3N4 with a high photodegradation activity toward methyl orange in a semiclosed system with a two-step heat treatment. By investigating the influence of heating temperature upon the thermal condensation of melamine, the optimal reaction condition for g-C3N4 was 520 °C for 2 h. Liu et al. [31] used urea as a precursor and produced g-C3N4 on a large scale by pyrolysis under ambient pressure without additive assistance. The retainable pyrolysis-generated self-supporting atmosphere and the reaction temperature are two necessary conditions.
Hydrothermal synthesis is also one of the most widely used methods, which is beneficial to control the accuracy in influencing reaction molar ratio and physio-chemical properties [1]. Wu et al. [32] prepared oxygen-containing-groups-modified g-C3N4 (OG/g-C3N4) through an in situ one-step hydrothermal treatment of bulk g-C3N4 in pure water. Hydrothermal treatment at 180 °C could promote the increase in the specific surface area of the resulting product from 2.3 to 69.8 m2 g−1, and oxygen-containing groups (-OH and C=O) were also successfully grafted on the surface of OG/g-C3N4 via the interlayer delamination and intralayer depolymerization. Due to its high surface area and oxygen-containing surface properties, OG/g-C3N4 demonstrated high photocatalytic performance on H2 evolution. Ahmad et al. [33] synthesized a highly efficient double Z-scheme g-C3N4/AgI/β-AgVO3 (g-CNAB) ternary nanocomposite using a one-pot hydrothermal route. They have characterized the optical properties, phase structure, and morphology of the as-prepared photocatalysts and evaluated their photocatalytic performance toward the photodegradation of different pollutants under visible-light irradiation. Experimental characterization indicated that g-CNAB possessed a dual Z-scheme heterojunction, which had the features of better spatial separation and charge-carrier transfer. As such, reactive species such as superoxide anion radical and hydroxyl radical can be favorably generated for the degradation of various contaminants.
To improve the photocatalytic performance of g-C3N4, templating strategy has also been applied to synthesize materials with unique appearance, structure, and properties. Due to the high specific surface area and low surface reflection, silica-based materials, including silica spheres, mesoporous silica, and SBA-15, usually serve as hard templates to fabricate g-C3N4. Sun et al. [34] used silica as a template to prepare a highly stable, hollow g-C3N4 nanosphere (HCNS) (Figure 2). Adjustable shell thickness performed as a light-harvesting platform for H2 evolution under visible light irradiation. However, corrosive reagents (e.g., NaOH, NH4HF2) are often used to remove silica templates, which are not friendly to the environment. On the contrary, the soft templating method uses ionic liquid as a template or co-template to fabricate g-C3N4, which is an environmentally friendly and one-step approach to prepare g-C3N4 with excellent performance. Zhao et al. [35] utilized cyanuric acid-melamine complex and an ionic liquid as soft templates to prepare hollow g-C3N4 spheres with a specific surface area as high as 84 m2 g−1. The morphology of g-C3N4 could be well controlled by adjusting ionic liquid and solvent. It has been demonstrated that the as-prepared hollow mesoporous carbon nitride exhibits ~30 times higher than traditional g-C3N4 in hydrogen production. Although significant progress has been achieved in the preparation and modification of g-C3N4, some preparation methods are neither environmentally friendly nor time-saving. Thus, it is necessary to develop green and facile synthesis routes.

3. Modification of g-C3N4

3.1. Morphology Control

Morphology control of g-C3N4 is one of the practical and effective methods to improve its photocatalytic performance. g-C3N4 can be regulated to different dimensions (Figure 3). Then, with g-C3N4 having a larger specific surface area, more adequate active sites can be obtained. Moreover, the visible-light response range of g-C3N4 will be expanded, and the carrier diffusion path will be shortened. This section will discuss 0–3D g-C3N4-based materials and their catalytic performance.

3.1.1. 0D g-C3N4

g-C3N4 dots are new members of the g-C3N4 family, smaller than 10 nm in size, and have quantum size effect, surface effect, and quantum confinement effect [36]. As the size effect causes the reverse motion of CB and VB, g-C3N4 with broad absorption (from ultraviolet to visible light) can be obtained. The photogenerated charge carriers favorably migrate to the particle surface for initiating oxidation or reduction reactions. In recent years, a variety of low-cost and size-controlled methods have been developed to prepare g-C3N4 dots with different physicochemical properties, including hydrothermal treatment [37], ultrasonic exfoliation [38,39], microwave-assisted solvothermal process [40], and solid reaction approach [1]. A hydrothermal and hot-air assisted chemical oxidation method was proposed to prepare g-C3N4 QDs by etching bulk g-C3N4 to graphene-like nanosheets [41]. Concentrated H2SO4 and HNO3 etched the nanosheets to produce g-C3N4 nanoribbons with sizes below 10 nm. Then, 5–9 nm g-C3N4 QDs can be obtained after the hydrothermal treatment of nanoribbons at 200 °C (pH = 5). The obtained g-C3N4 QDs exhibited a strong blue emission and upconversion behavior, promising visible-light-driven metal-free photocatalytic systems. Liu et al. [38] synthesized g-C3N4 QDs by recrystallization and ultrasonic exfoliation from the precursor of dicyandiamide. g-C3N4 QDs with different sizes (5–200 nm) can be prepared by adjusting ultrasonic time up to 90 min. Chen et al. [40] prepared 0D/2D CNQDs/g-C3N4 isotype heterojunctions by a simple microwave assisted-polymerization method. The obtained product exhibited excellent photocatalytic performance toward norfloxacin degradation, and its reaction rate was as much as two times higher than pristine g-C3N4. Zero-dimensional QDs structure materials with nanometer size have a large surface/volume ratio, abundant surface atoms, and unsaturated coordination state, which conduce to their high activity for photocatalysis [42,43]. In addition, the absorption spectra of QDs/g-C3N4 will appear blue shift, and band gaps will be broadened, affecting their electronic band structures in photocatalysis [44]. Thus, grafting 0D QDs on g-C3N4 is a promising approach to creating great reactive active sites and enhancing photoelectric conversion ability to improve photocatalytic activity.

3.1.2. 1D g-C3N4

As morphology significantly affects materials’ photochemical properties and electron transfer rate, the photocatalytic activity of g-C3N4 can be improved by adjusting its size and shape. The polymer characteristics of g-C3N4 make it an excellent flexible structure. One-dimensional g-C3N4, including nanowires [45], nanorods [46,47], and nanotubes [48], can be prepared by non-metallic hard-template, soft-template, self-template, and template-free methods. Bai et al. [47] found it possible to transform g-C3N4 from nanoplates to nanorods in a simple reflux way. They compared the photocatalytic activity and intensity of both shapes; results achieved in their study demonstrated that the photocatalytic activity and power of nanorods were ~1.5 and 2.0 times higher than those of nanoplates under visible light, attributable to an increase in active lattice face and elimination of surface defects. Jiang et al. prepared melamine crystals by a transitional metal derived re-crystalline process and then created g-C3N4 nanotubes with melamine crystals through a thermal polymerizing reaction method. They have applied transitional metal ions (Fe3+, Co2+, Ni2+, and Mn2+) in the growing of melamine crystals and have characterized the obtained ion-modified g-C3N4 nanotubes with XRD, FT-IR spectra, and XPS (Figure 4). It has been demonstrated that Fe3+-ion-modified g-C3N4 nanotubes (Fe3+ R-650 CN) exhibited enhanced absorbance at 500 nm and decreased band gap. The hydrogen evolution rate (7538 μmol h−1 g−1) is almost 13.5-fold than that of conventional g-C3N4 nanosheets. Mo et al. [48] synthesized defect-engineered g-C3N4 nanotubes through an efficient self-assembled method, and applied them to hydrogen evolution. Around 6.8% external quantum efficiency was achieved at 420 nm. Among 1D g-C3N4, nanotubes are superior to others because they allow more effective light absorption, offer more active sites, and invent different electron pathways with tube morphology [25]. Generally, due to the nanometer scale of 1D g-C3N4-based materials in the radial direction, the diffusion distance (from volume to the surface) of photoexcited charges would be reduced, and the charge separation during photocatalytic reactions would be promoted. In addition, if g-C3N4-based materials are transformed from 2D to 1D structure, polygonal defects may appear to form the active sites and increase the contact surface of reactions, thus improving the photocatalytic performance.

3.1.3. 2D g-C3N4

Bulk g-C3N4, prepared by the traditional one-step calcination method, has a small specific surface area (10 m2 g−1), and the photogenerated electrons and holes tend to recombine, which reduces its photocatalytic activity [50]. Since g-C3N4 has a 2D layered structure connected by van der Waals force, it will produce unique chemical and physical properties as below when it is stripped into multilayer or single-layer nanosheets. First, the large surface area and easy access to active reaction centers facilitate interactions with reactants. Second, nanoscale thickness or even less can minimize the migration distance of charge carriers, ensuring the rapid transport of charge carriers from the bulk phase to the surface of a catalyst. As a result, it would effectively inhibit electron-hole pairs from recombination. Third, the unique two-dimensional flexible planar structure can enhance compatibility with various modification strategies, such as heterojunction construction, cocatalyst modification, and vacancy introduction. This feature further improves the quantum efficiency in a photocatalysis process [51]. Various methods have been developed to prepare g-C3N4 nanosheets, such as thermal oxidation etching, chemical peel etching, supramolecular self-assembly, and ultrasonic treatment of exudation [1,51]. Among these, thermal oxidation etching is the most common method, which can overcome van der Waals force between layers and peel bulk g-C3N4 into 2D nanosheets under high-temperature oxidation conditions. Niu et al. [52] prepared g-C3N4 nanosheets (~2 nm thickness) by thermal oxidation etching of bulk g-C3N4 in the air. UV-visible absorption (Figure 5A) exhibits a blue shift of the intrinsic absorption edge in the nanosheets. Compared to the bulk (2.77 eV), the nanosheets possess a higher specific surface area and 0.2 eV larger bandgap (Figure 5B). These characteristics benefit electron transport along the in-plane direction and increase the lifetime of photoexcited charge carriers because of the quantum confinement effect. Furthermore, the average hydrogen evolution rate of nanosheets is 170.5 μmol h−1 under UV-visible light, which is 5.4 times higher than that of the bulk counterpart. Due to the simplicity of operation, the supramolecular self-assembly method has attracted much attention. For example, conventional melamine-cyanuric acid (MCA) complexes can be obtained by mixing melamine and cyanuric acid in a solution. However, the use of solvent limits the batch preparation of MCA complexes. To address this issue, Liu et al. [53] proposed synthesizing supramolecular precursors through hydrothermal treatment of dicyandiamide and prepared 3D holey g-C3N4 nanosheets with excellent photocatalytic performance. The precursor exhibited a similar structure to that of conventional MCA. In contrast, their thermal decomposition and morphology were different, which led to the distinction of microstructures, optical properties, charge recombination, photoelectrochemical behavior, and photocatalytic activity. The holey g-C3N4 can make up for the shortcomings of recombination of charge carriers, retarded visible light utilization, and the limited surface-active sites in the bulk g-C3N4 catalysts, which could contribute to their outstanding application in photocatalytic hydrogen evolution.

3.1.4. 3D g-C3N4

Three-dimensional nanostructures are considered an effective way to improve the properties of photocatalytic materials because they can provide a larger specific surface area and more reactive active sites. Although soft and hard template methods have been widely used in the preparation of 3D g-C3N4, they generally employ toxic substances and organic solvents to remove the template, which limits their application in actual green production. Salt templates have replaced some traditional templates (silica, sodium dodecyl sulfate, anodic alumina) for safe synthesis processes. Qian et al. [54] introduced a simple and effective sodium-chloride-assisted ball milling method to prepare 3D porous g-C3N4. Three-dimensional cubic sodium chloride particles can be used as an easily removable template to design 3D porous structures and as a limiting structure to prevent aggregation of g-C3N4 during calcination. The modified 3D interconnecting network structure of g-C3N4 has a large specific surface area, significantly improving the photocatalytic performance. The hydrogen production rate can be as high as 598 μmol g−1 h−1 with 3.31% quantum efficiency at 420 nm. In addition, ionic liquids are widely used in many fields due to their excellent fluidity and solubility. In the process of nanomaterial preparation, the ionic liquid can self-assemble into micelles, which impacts the size and morphology of the nanomaterials. For example, Zhao et al. [35] controlled the morphology of hollow mesoporous g-C3N4 spheres by changing ionic liquid concentration. At a low ionic liquid concentration, the prepared mesoporous g-C3N4 showed a hollow spherical structure while, at a high ionic liquid concentration, the cyanate-melamine (CM) complex rearranged, which could induce the formation of a flower-like structure with ultrathin nanosheet (Figure 6). Moreover, the hollow as-prepared g-C3N4 exhibits higher light absorption in the visible range and a faster separation rate of photogenerated hole-electron pairs than bulk C3N4. In addition, the hydrogen production of as-prepared hollow mesoporous g-C3N4 exhibits ~30 times higher than traditional g-C3N4, because of its high surface area.
In addition, some unique g-C3N4-based structures, like the “seaweed” network, spiral rod, and hollow fusiform, have also been reported [55,56,57,58,59], and a superior photocatalytic activity to bulk samples was observed for higher specific surface areas. Among these morphologies of g-C3N4, 1D nanoribbons, with a large specific surface area, more active sites, a short diffusion distance, and preferential directions for photo-generated electron-hole carriers, effectively accelerate the catalytic performance of g-C3N4. Better yet, the 3D morphology assembled from 1D units have depressed agglomeration, increased exposure of active sites, and decreased mass transportation resistance.

3.2. Elemental Doping

Heteroatom doping effectively regulates the electrical, optical, and structural properties of semiconductors by introducing active impurities [60]. In general, non-metallic heteroatoms can participate in the C3N4 lattice and partially replace C or N atoms, while metal atoms can insert into the triangular gap cavity of the g-C3N4 lattice. Doped metal and non-metal atoms generate intermediate band gaps near CB and VB to regulate the band structure of g-C3N4, which can effectively realize the separation and transmission of electron-hole pairs and broaden their optical response range [55,60]. Therefore, doping is a prevalent method to improve the photocatalytic performance of semiconductors.

3.2.1. Non-Metal Doping

Non-metal element doping could maintain the metal-free character of g-C3N4. Additionally, due to the high ionization energies and electronegativity of non-metals, they can quickly form covalent bonds with other compounds by gaining electrons during the reaction process [61,62]. The introduction of non-metals will break the symmetry of g-C3N4 and result in a faster separation speed of electron-hole pairs [63]. The ordinary non-metallic doping atoms include O, P, S, B, C, N, and halogens (F, Cl, Br, and I). Among them, the O atom, one of the most typical non-metallic doping elements, has shown extraordinary potential in improving the photocatalytic performance of g-C3N4. Zhang et al. [64] presented a hydrothermal method and fabricated a porous and oxygen-doped g-C3N4 photocatalyst for efficient photocatalytic hydrogen production by forming homogeneous supramolecular complexes (Figure 7). They introduced porous structure and heteroatom doping in g-C3N4 to adjust its active sites and electronic structure for enhanced light harvesting, charge separation, and transfer. Compared with bulk g-C3N4, the hydrogen evolution activity of the g-C3N4 photocatalysts is 11.3-fold higher than bulk g-C3N4. She et al. [65] introduced oxygen in g-C3N4 and prepared 2D porous ultrathin oxygen-doped g-C3N4 nanosheets. It has been demonstrated that the band gap was enlarged (~0.20 eV), and the transport ability of photogenerated electrons and the redox ability were improved, which is caused by the quantum confinement effect. Besides, the specific surface area of non-metal doped g-C3N4 is larger (~20 times) than that of the bulk, which will supply more active sites with adequate quality and offer more adsorption sites. In short, due to the increased bandgap, the introduction of the electrophilic groups and the morphology structure, the electron-hole recombination probability is inhibited and the redox ability will be improved, which contribute to the enhanced photocatalytic activity.

3.2.2. Metal Doping

In terms of metal-doped g-C3N4, the porous structure of heptamine and electron-rich sp2 nitrogen atoms can provide sites for metal coordination, and doped metal can easily bind to the three neighboring N atoms in the form of g-C3N4. Metal element doping, especially alkali, has been demonstrated to reduce the energy gap, supply more active reaction sites, adjust VB position and improve photocarrier separation to enhance the visible light absorption [66]. Due to the uneven distribution of semiconductor charge space, the improvement of photogenerated carrier separation efficiency is increased. In addition, the alkali metal doping can increase the π-conjugated systems and reduce the recombination rate of the semiconductor electron-hole pairs, which will be beneficial to improve the efficiency of H2 evolution by photolysis [67]. The ordinary doping metal atoms are K, Na, Ag, Au, Fe, Ni, Pt, etc. Gao et al. [68] designed a simple one-step pyrolysis process to synthesize Fe-doped g-C3N4 nanosheets with NH4Cl as a “dynamic gas template” and FeCl3 as a Fe source, respectively (Figure 8). The experimental results show that Fe species may exist at the state of Fe3+ and form Fe-N bonds in g-C3N4, thereby expanding visible light absorption regions and reducing the band gap of g-C3N4 nanosheets. Moreover, doping specific amounts of Fe could promote exfoliation and increase the specific surface area of g-C3N4, while excessive Fe might break the sheeting structure. The specific surface area of optimized Fe-doped g-C3N4 nanosheets reached 236.52 m2 g−1, which was 2.5 times higher than g-C3N4 nanosheets. In addition, Deng et al. [69] prepared K+ and cyano-group co-doped crystalline polymeric carbon nitride (KC-CCN) by a one-step thermo-polymerization approach. They applied thiourea and potassium thiocyanate as precursors, and the resulting KC-CCN demonstrated a highly crystal structure, stronger light harvesting, and a higher electron-hole separation ratio.

3.3. Heterojunction Construction

The key factor that restricts the activity of semiconductor photocatalysts is improving the separation and transport efficiency of electron-hole pairs. Thus, the construction of g-C3N4-based heterojunction photocatalysts is one of the most common and effective methods. When two semiconductors with different band structures combine to form a heterojunction, effective charge transfer will be included at the interface, improving charge separation and transport efficiency [70,71,72,73,74]. In addition, the light absorption capacity of a photocatalytic system can be enhanced by combining it with narrow-band gap semiconductors. For example, Xu et al. [75] proposed a wet chemical method to fabricate CdS/g-C3N4 (CSCN) heterojunctions in situ. Through the result of XRD, FTIR, TEM, and optical band gap for CSCN, the formation of heterojunctions was confirmed. The CdS nanoparticles dispersed uniformly on the surface of g-C3N4 nanosheets and the interfaces between g-C3N4 and CdS in composites is very close, which can efficiently enhance the electron transfer between the two semiconductors. Comparing the UV–vis DRS spectra of g-C3N4 (474.9 nm) and CdS (685.7 nm), the absorption thresholds of all CSCN composites locate between that of g-C3N4 and CdS, which indicates strong visible-light absorption. Among the materials, including the individual CdS, g-C3N4, and different CSCN composites, CSCN733 possesses the highest adsorption capacity and exhibits the highest methyl orange degradation efficiency, 100% with 40 min adsorption. Liu et al. [76] embedded nanorod-like CoP nanoparticles into g-C3N4 nanosheets to form CoP-CN heterostructure. The XRD data indicated that the 0.5% CoP-CN hybrid incorporates the representative peaks of g-C3N4 and CoP with g-C3N4 demonstrating the main phase. The TEM of 0.5% CoP-CN composites displays a porous and fluffy structure. The binding energy of P 2p3/2 is lower than that of P 2p3/2, while the binding energy of Co 2p3/2 eV is a little higher than that of metallic Co 2p3/2. This result indicates that the electron transfer from Co to P to form Co-P covalent bonds results in a small positive charge of Co and negative charge of P. This finding would account for the excellent activity of CoP-based photocatalysts in the HER process, in which Co serves as active center while P performs as the proton acceptor. The flat band potentials of 0.5% CoP-CN were decreased to −0.28 V and the CB tuned upward for more negative potential, achieving more efficient interfacial charge transportation and separation by establishing a certain inner electric field. Furthermore, among the evaluation of photocatalytic water half-splitting for H2 production, 0.5% CoP-CN exhibit excellent activity and reached 959.4 μmol h−1 g−1, which is almost 3.1-fold than that of pristine g-C3N4 nanosheets. This can be attributed to its decreased over-potentials, more negative photo-reductive potentials, increased interfacial charge transfer efficiency, and higher solar-to-hydrogen efficiency. Ma et al. [77] combined TiO2 with g-C3N4 to form a Z-type heterojunction, which can effectively separate photogenerated electrons and holes and improve the photocatalytic activity. However, the utilization of visible light of TiO2 is very low because of its large band gap. Thus, to improve this weakness, the Eg of TiO2 should be reduced, followed by its potential change of valence band and the conduction band. Such a move would change the heterojunction type, which is not consistent with improving the catalytic efficiency of composite materials to visible light. Therefore, various kinds of g-C3N4-based heterojunction photocatalysts have emerged to enhance the efficiency of photogenerated carrier separation further and light absorption capacity, including ternary [78,79,80,81,82,83] and other type II [84,85,86,87,88] and S-type heterojunction [89,90], Schottky junction [91,92], and van der Waals heterojunction [93,94].
The S-type heterojunction is composed of reduction photocatalysts (RPs) and oxidation photocatalysts (OPs) with staggered band structures, similar to the type-II heterojunction but with an entirely different charge-transfer route [95]. S-type photocatalysis is an effective way to control charge separation in various photocatalytic reactions. Since Yu et al. first proposed the concept of S-type heterojunction in 2019 [96], a large number of studies on the g-C3N4-based S-type heterojunction have been reported, such as Sb2WO6/g-C3N4 [97], S-doped g-C3N4/TiO2 [98], ZnFe2O4/g-C3N4 [99], g-C3N4/Zn0.2Cd0.8S-DETA [100], g-C3N4/Bi/BiVO4 [101], WO3/g-C3N4 [102], etc. The Schottky junction photocatalyst has also attracted much attention. Generally, conductor-semiconductor heterojunction has two main combined modes: Schottky junction and ohmic contact. Conductors and semiconductors with different Fermi levels will generate Schottky effects at their contact interfaces to induce internal electric fields to drive charge flow until the system reaches equilibrium. The proper orientation of built-in electric fields will promote the directional separation of charge carriers, leading to the practical generation of photogenerated charge carriers and improving the photocatalytic activity. A large number of g-C3N4 Schottky junction photocatalysts have been reported, such as CoP/g-C3N4 [103], CuS/g-C3N4 [104], Ti3C2/g-C3N4 [105], carbon/g-C3N4 [106], MoO2/g-C3N4 [107], Cu-NPs/g-C3N4 [108], etc. Recently, van der Waals (vdW) heterojunction has been proposed to regulate the electrical and optical properties of 2D materials accurately. vdW heterojunction not only overcomes the lattice matching limitation for enhancing interfacial charge separation and transfer but also leads to strong electronic coupling between layers to improve catalytic activity [109]. The research on g-C3N4-based vdW heterojunctions is currently enjoying a boom, and different types have been reported, including phosphorene/g-C3N4 [110], g-C3N4/Zn-Ti LDH [111], g-C3N4/C-doped BN [112], g-C3N4/COF package-TD [113], etc.
Moreover, metal-organic framework (MOF) materials have exhibited excellent photocatalytic performances due to their unique porous structures and favorable transfer of e and h+ [114,115,116]. However, most MOFs have low stability and weak light response. Therefore, the heterojunction of MOF and g-C3N4 materials has become popular in recent years. For example, Zhang et al. [117] synthesized a novel hybrid of Zr-based metal-organic framework with g-C3N4 (UiO-66/g-C3N4) nanosheets (10:10) by annealing their mixture. The photoelectron can transfer efficiently from the CB of g-C3N4 to that of UiO-66 through the inner electric field generated by the heterojunction, which is beneficial to decrease the recombination of electron/hole. Together with their porous structures, much more organic dye molecules can absorb on the surface of the heterojunction catalyst, thus facilitating the electron/hole transfer and enhanced photocatalytic activity. Han et al. [114] prepared TPVT (tridentate ligand 2,4,6-tris(2-(pyridin-4-yl)vinyl)-1,3,5-triazine)-MOFs and combined them with g-C3N4. It has been demonstrated that the TPVT-MOFs@g-C3N4-10 can reach 56.4 μmol·g−1·h−1 in CO2 reduction, which is 3.2-fold higher than that of g-C3N4. All these researches have provided a new insight into the design of g-C3N4 based photocatalysts to deal with the organic dyes in environment.

3.4. Modification with Carbon Nanocomposites

The weak van der Waals forces between layers and the abundant hydrogen bonds in the molecular structure make g-C3N4 exhibit slow charge transfer kinetics and poor electrical conductivity. Carbon materials have been widely used in photocatalysis due to their low price, good conductivity, high stability, non-toxicity, and harmlessness. Many composite photocatalysts with excellent photocatalytic performance have been prepared by combining carbon nanomaterials (such as fullerene [118], graphene [119], carbon nanotubes [120], etc.) with g-C3N4, which have unique nanostructures and excellent electron-optical properties (Figure 9). The introduction of carbon material reduces the electron-hole pair recombination rate of photocatalysts and improves the photo absorption, thus improving the photocatalytic performance of g-C3N4-based materials. g-C3N4 photocatalysts modified by carbon materials can promote photocatalytic reaction through heterojunction interaction, cocatalyst effect, surface recombination, local charge modification, and other ways [121]. For example, Yuan et al. [122] prepared a graphene-g-C3N4 composite photocatalyst by calcining graphene with melamine, and excellent photocatalytic degradation performance toward RhB was observed under acidic conditions. Ge et al. [123] prepared multi-walled carbon nanotubes (MWNTs)/g-C3N4 composite photocatalyst by heating MWNTs and g-C3N4, in which MWNTs favored the efficient separation of photo-generated charge carriers. As a result, this material exhibited unique performance in photocatalytic H2 production under visible light conditions.
As a new carbon-based nanomaterial, carbon dots (CDs) exhibit excellent up-conversion photoluminescence and remarkable photogenerated charge-carrier transfer and reservoir [124,125,126]. They can also modify g-C3N4 and broaden optical absorption by reducing the electron-hole pair recombination rate [127,128]. Such features have caused extensive attention to CDs-modified g-C3N4. For instance, Fang et al. [129] prepared a CDs-modified g-C3N4 hybrid by dicyandiamide and CDs obtained from the combustion soot of alcohol. Based on the investigation of CDs modification on the structure and photocatalytic activity of g-C3N4, they found that CDs modification caused the lattice distortion of g-C3N4, and CDs performed as an electron sink, which could prevent the recombination of photo-generated electron-hole pairs. Ai et al. [130] reviewed the combination methods of g-C3N4 and CDs (Figure 10) for enhancing photocatalytic performance and indicated that g-C3N4/CDs hybridization has strong practicability in efficient photocatalytic hydrogen generation, photocatalytic carbon dioxide reduction, and organic pollutant degradation. However, it is still in the early stage. Much effort should be made to develop green and facile synthesis routes and solve the insufficient utilization of visible and near-infrared light.

4. Photocatalytic Application in Energy and Environmental Sustainability

The energy and environmental crises have been an ongoing challenge, which is related to the vital interests of people around the globe. How to solve this problem through sustainable development strategies is considered deeply by scientific researchers. Photocatalysis provides a powerful technique for fully utilizing solar in the field of energy conversion [28,131]. Here, we will mainly introduce the photocatalytic application of g-C3N4 in energy and environmental remediation, including H2 production, CO2 photoreduction, and pollutant degradation.

4.1. H2 Production

Hydrogen is gathering strong momentum as a pivotal energy transition pillar driven by the global shift toward decarbonization. Nevertheless, 85% of H2 is produced from fossil fuel combustion, which generates roughly 500 metric tons of carbon dioxide every year and proffers a challenge and obstacle toward the sustainable living of future generations [132]. Solar-driven photocatalytic H2 generation as a promising technology has received extensive attention in addressing the global energy crisis [133,134]. Photocatalytic water splitting for the energy transformation from solar to eco-friendly fuels has been studied for decades with various semiconductor photocatalysts. As a type of semiconductor photocatalyst, g-C3N4 is simple and inexpensive to fabricate, and has an adequate bandgap (≈2.7 eV) for activation upon sunlight irradiation. Wang’s group first utilized g-C3N4 in photocatalytic H2 evolution [15,135]. Nonetheless, pristine g-C3N4 is far from satisfactory energy conversion because of its low light energy utilization, low density active sites, and ineffective isolation of the photogenerated excitons. Thus, researchers have proposed numerous strategies to boost the photocatalytic activity of g-C3N4-based materials for H2 production. For example, the g-C3N4/carbon-dot-based nanocomposites, which possess enormous visible light absorption and applicable energy structures, have been prepared and serve as efficacious photocatalysts in photocatalytic water splitting for H2 generation under light illumination [128,134,136,137]. Gao et al. reported hexagonal tubular g-C3N4/CD-based nanocomposites which exhibited nine times higher than bulk g-C3N4 in H2 production rate [134] and related results indicated that CDs performed as both photosensitizer and electron acceptor. CDs could absorb long wavelength light to extend the visible-light response region and suppress the recombination of electron-hole pairs. Hussien et al. [138] combined four different strategies (non-metal doping, porosity generation, functionalization with amino groups, and thermal oxidation etching) in a one-pot thermal reaction and successfully prepared amino-functionalized ultrathin nanoporous B-doped g-C3N4 by using NH4Cl as a gas bubble template, together with a thermal exfoliation process to produce ultrathin sheets (Figure 11). According to the process, the surface area, adsorption capacity, and charge migration of the as-prepared photocatalyst have been improved, and a 3800 µmol g−1 h−1 H2 generation rate and 10.6% prominent quantum yield were recorded. Li et al. [139] decorated carbon self-doping g-C3N4 nanosheets with gold-platinum (AuPt) nanocrystals through a photo-deposition route and compared the photocatalytic H2 evolution performance of Pt/CCN, Au/CCN, Au/Pt/CCN, and Pt/Au/CCN, in which AuPt/CCN stood out and gave the highest H2 generation rate (1135 μmol h−1). The excellent performance can be ascribed to the non-plasmon-related synergistic effect of Au and Pt atoms in AuPt nanocrystals. Sun et al. [140] assessed the arrangements of metal- and non-metal-modified g-C3N4 composites in hydrogen evolution and found that the contribution of dye conjugation in non-metallic g-C3N4 composites favored their performance (Figure 12). However, the co-catalyst doping strategy was recommended for metallic g-C3N4 composites. In addition, the hybrid of MOF materials and g-C3N4 is also a good approach to develop novel photocatalysts. For example, Devarayapalli et al. [141] reported a g-C3N4/ZIF-67 nanocomposite and obtained a 2084 μmol g−1 H2 production, which is 3.84-fold greater than that of bare g-C3N4.
Based on the descriptions mentioned above, Table 1 compares the performance of different g-C3N4-based materials for photocatalytic H2 generation reported within the last three years.

4.2. CO2 Photoreduction over g-C3N4

Rising atmospheric levels of CO2 and the consumption of fossil fuels raise a concern about the continued reliance on the utilization of fossil fuels for both energy and chemical production [173]. Photocatalytic reduction of CO2 is a promising strategy to meet increasing energy needs and reduce the greenhouse effect [174]. Through photocatalytic reduction, CO2 can be converted to light oxygenates and hydrocarbons. Photocatalytic CO2 reduction is a multielectron transfer process. Fu et al. [175] have listed the possible reaction and corresponding redox potentials and stated that CO2 was complicated to reduce at room temperature due to its stable chemical structure. For the complex reaction, five factors, comprising the matching of band energy, separation of charge carrier, kinetic of e- and hole transfer to CO2 and reductant, the basicity of photocatalyst, and the strength and coverage of CO2 adsorption, are considered to be crucial [176]. As a hot member of photocatalysts, g-C3N4 has been applied to CO2 photo-reduction in recent years because the CB of g-C3N4 is sufficient to reduce CO2 to various hydrocarbons, such as CH3OH, CH4, HCHO, and HCOOH, etc. (Figure 13) [177].
However, metal-free g-C3N4 is limited for CO2 reduction activity due to its poor ability to activate the C-O bond of CO2. To improve the photocatalytic movement of CO2 conversion, different metal units have been composited with g-C3N4 for broadening the absorption response range, and accelerating the charge separation and transfer, such as Pt/g-C3N4 [178], Co2+/g-C3N4 [179,180], Au/g-C3N4 [181], etc. Metal nanoparticles acting as cocatalysts could effectively improve the photocatalytic activity and selectivity of CO2 reduction. In addition, other methods, including doping, loading cocatalysts and nanocarbons, constructing Z-scheme, and heterojunction, have also been employed [16,182,183,184,185,186,187,188,189]. For example, Fu et al. [190] prepared hierarchical porous O-doped g-C3N4 nanotubes (OCN-Tube) through continuing thermal oxidation exfoliation and curling condensation of bulk g-C3N4. Due to the higher specific surface area, better light harvesting, higher CO2 uptake capacity, and superior separation efficiency of photogenerated charge carriers, the OCN-Tube exhibits excellent photocatalytic CO2 reduction performance into CH3OH. The CH3OH evolution rate was as high as 0.88 µmol g−1 h−1, five times higher than the bulk (0.17 µmol g−1 h−1). Huo et al. [191] fabricated amine-modified step-scheme (S-scheme) porous g-C3N4/CdSe-diethylenetriamine (A-PCN/CdSe-DETA) by a one-step microwave hydrothermal method. The modification by amine and formation of S-scheme heterojunction contributed to the remarkable photocatalytic performance of A-PCN/CdSe-DETA composite in CO2 reduction and a CO production rate of 25.87 μmol/(h g) was achieved under visible-light irradiation. Wang et al. [174] reviewed different modification methods of g-C3N4-based photocatalysts for CO2 reduction. They discussed each method (including morphology adjustment, co-catalysts, heterostructures, and doping) and compared the theoretical calculations and experimental results. By morphology adjustment, g-C3N4 with various shapes can be fabricated, such as rods, tubes, nanosheets, hollow spheres, and honeycomb-like structures. Due to the advantage of cocatalysts (e.g., Au, Ag, Pt, Pd, MXene, AuCu alloy, Pd-Ag), g-C3N4 with co-catalysts can be widely applied to activate CO2 on the surface. Heterojunction with different types is also an effective method to improve the properties of g-C3N4-based materials. In addition, elemental doping is considered a common method to enhance photocatalytic quantum efficiency by changing the energy band, surface electronic property, and electrical conductivity. Table 2 compares the performance of different g-C3N4-based materials for photocatalytic CO2 reduction reported within the last three years.

4.3. Degradation of Organic Pollutants

Along with rapid population growth and significant industrialization development, large numbers of toxic, hazardous, and endless contaminants invade the environment, threatening to human life, especially a variety of pollutants present in water that are difficult to eliminate or degrade naturally. Photocatalytic degradation of contaminants is a green and efficient technology for coping with sewage [128,223]. Different kinds of g-C3N4-based materials (Table 3) have been exploited to increase the photodecomposition efficiency of pollutants, such as the constructed heterojunction, loading O2-reduction co-catalysts, g-C3N4/CDs-based nanocomposites, and so on [182,224,225,226]. Generally, under the irradiation of visible light, the photogenerated electrons (e) on the g-C3N4 catalyst will be excited from VB to CB, leaving holes (h+) in the VB. The holes can oxidize pollutants directly or react with H2O/OH to form hydroxyl radicals [227]. When the REDOX potential of g-C3N4 composites is more negative than O2/O2, the photogenerated electrons in the material can react with O2 to produce O2 with strong oxidation capacity [228]. In addition, the resulting O2 could be protonated to produce OH [229]. Finally, the RhB dye is degraded to CO2 and H2O under the action of these free radicals (Figure 14). Chen et al. [230] fabricated a BiFeO3/g-C3N4 heterostructure through mixing-calcining and compared its performance with BiFeO3. Around 30% higher photocatalytic efficiency toward RhB dye was observed for the BiFeO3/10% g-C3N4 heterostructure, which was assigned to the contribution of a higher concentration of O2. Zhang et al. [231] studied the selective reduction of molecular oxygen on g-C3N4 and probed its effect on the photocatalytic phenol degradation process. Compared with bulk g-C3N4, the exfoliated nanosheet yielded a three times improvement in photocatalytic phenol degradation. It has been demonstrated that bulk g-C3N4 prefers to reduce O2 to O2via one-electron reduction. At the same time, the photoexcited g-C3N4 nanosheet facilitates the two-electron reduction of O2 to yield H2O2 because of the formation of 1,4-endoperoxide species. The two-electron reduction of O2 on the nanosheet surface boosts hole generation and thus accelerates phenol oxidation degradation [231,232]. Thus, to improve the photocatalytic performance of g-C3N4, more effort should be devoted to strengthening the solid O2-reduction reactions. For example, Liu et al. [83] reported a heterojunction material of K-doped g-C3N4 nanosheet -CdS and degraded tetracycline with 94% degradation under visible light in 30 min. In addition, due to the electronegativities, ionic radius differences, and impurity states, element doping is also an effective method to manipulate the electronic structure and physicochemical performance of g-C3N4-based materials. Gao et al. [68] synthesized Fe-doped g-C3N4 nanosheets and obtained 1.4- and 1.7-fold higher degradation rates of MB than that of pure g-C3N4 nanosheets and bulk g-C3N4, which indicated that the exploitation of efficient g-C3N4-based photocatalysts with high stabilization and degradation under visible light irradiation would significantly contribute to sewage disposal. Zhang et al. [117] synthesized a novel hybrid of Zr-based metal-organic framework with g-C3N4 (UiO-66/g-C3N4) nanosheets and applied a photodegradation of methylene blue, by which a 100% photodegradation was achieved within 4 h under visible light. This research has provided a new insight into the design of g-C3N4-based photocatalysts to deal with organic dyes in the environment.

5. Conclusions and Future Perspective

g-C3N4-based materials are still a research hotspot in photocatalysis, especially their application in energy and environmental sustainability. Although significant progress has been achieved in the preparation and modification of g-C3N4, several issues remain to be resolved in future research: (1) Some preparation methods are neither environmentally friendly nor time-saving. Thus, it is necessary to develop green and facile synthesis routes. For example, it should be encouraged to use plant leaves, natural halloysite, and some natural raw materials in the preparation of g-C3N4-based materials. (2) The absorption ability of available g-C3N4-based materials to visible and near-infrared light is still low, which is not beneficial to improving solar energy utilization. Coupling g-C3N4 with visible and near-infrared CDs might be an effective strategy. It would efficaciously improve the e/h+ pair separation capability and visible light harnessing capability, thus enhancing the related photocatalytic performance. (3) Some structures of modified g-C3N4-base materials are complex, and the corresponding photocatalytic reaction mechanisms is not clear yet. Introducing density functional theory could provide insights into the photocatalytic mechanisms via disclosing the materials’ structural, electronic, optical, and other properties. Detailed reaction processes can be performed by using in situ monitoring techniques (e.g., in situ infrared spectroscopy and mass spectrometry) to capture the reactive intermediates. (4) Although microscopic techniques and time-resolved spectroscopy have achieved the study of the steady-state charge distribution and charge transfer dynamics of photocatalysts, tracking the spatiotemporally evolving charge transfer processes in single photocatalyst particles and elaborating their exact mechanism is still a great challenge. Thus, it is significant to develop techniques to map holistic charge transfer processes at the single-particle level, identify where charges go and reveal how long they live on different sites. (5) Finally, the integration of artificial intelligence (AI) and other interdisciplinary techniques will play a tremendous driving role in precisely designing g-C3N4-based photocatalysts with excellent performance. For example, AI models could be developed to correlate photocatalytic performance with experimental conditions, which may help predict the photocatalytic performance of g-C3N4-based materials, improve the trial-and-error paradigm, and design new composite structures.

Author Contributions

Conceptualization of the review, Q.W. and F.H.; writing—original draft preparation, Q.W., Y.L. and S.S.; writing—review and editing, G.A., X.X. and B.Z.; supervision, Y.Z. and Z.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Basic Research Program of Shaanxi Province of China (Grant No. 2021GY-247, 2019JC-33, and 2023-JC-QN-0122), National Natural Science Foundation of China (No. 22274128), the Youth Innovation Team of Shaanxi Universities (Grant No. Z19257), and China Postdoctoral Science Foundation (Grant No. 2022MD723840).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Li, W.; Sohail, M.; Anwar, U.; Taha, T.; Al-Sehemi, A.G.; Muhammad, S.; Al-Ghamdi, A.A.; Amin, M.A.; Palamanit, A.; Ullah, S. Recent progress in g–C3N4–based materials for remarkable photocatalytic sustainable energy. Int. J. Hydrogen Energy 2022, 47, 21067–21118. [Google Scholar] [CrossRef]
  2. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. Dual cocatalysts in TiO2 photocatalysis. Adv. Mater. 2019, 31, 1807660. [Google Scholar] [CrossRef] [PubMed]
  3. Xinchen, W.; Maeda, K.; Thomas, A.; Takanabe, K.; Gang, X.; Carlsson, J.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar]
  4. Zhu, H.; Cai, S.; Liao, G.; Gao, Z.F.; Min, X.; Huang, Y.; Jin, S.; Xia, F. Recent advances in photocatalysis based on bioinspired superwettabilities. ACS Catal. 2021, 11, 14751–14771. [Google Scholar] [CrossRef]
  5. Xing, Y.; Wang, X.; Hao, S.; Zhang, X.; Wang, X.; Ma, W.; Zhao, G.; Xu, X. Recent advances in the improvement of g-C3N4 based photocatalytic materials. Chin. Chem. Lett. 2021, 32, 13–20. [Google Scholar] [CrossRef]
  6. Ismael, M. A review on graphitic carbon nitride (g-C3N4) based nanocomposites: Synthesis, categories, and their application in photocatalysis. J. Alloys Compd. 2020, 846, 156446. [Google Scholar] [CrossRef]
  7. Meng, A.; Wu, S.; Cheng, B.; Yu, J.; Xu, J. Hierarchical TiO2/Ni (OH)2 composite fibers with enhanced photocatalytic CO2 reduction performance. J. Mater. Chem. A 2018, 6, 4729–4736. [Google Scholar] [CrossRef]
  8. Li, Y.; Zhou, M.; Cheng, B.; Shao, Y. Recent advances in g-C3N4-based heterojunction photocatalysts. J. Mater. Sci. Technol. 2020, 56, 1–17. [Google Scholar] [CrossRef]
  9. Patnaik, S.; Sahoo, D.P.; Parida, K. Recent advances in anion doped g-C3N4 photocatalysts: A review. Carbon 2021, 172, 682–711. [Google Scholar] [CrossRef]
  10. Govindasamy, P.; Kandasamy, B.; Thangavelu, P.; Barathi, S.; Thandavarayan, M.; Shkir, M.; Lee, J. Biowaste derived hydroxyapatite embedded on two-dimensional g-C3N4 nanosheets for degradation of hazardous dye and pharmacological drug via Z-scheme charge transfer. Sci. Rep. 2022, 12, 11572. [Google Scholar] [CrossRef]
  11. Ye, L.; Liu, J.; Jiang, Z.; Peng, T.; Zan, L. Facets coupling of BiOBr-g-C3N4 composite photocatalyst for enhanced visible-light-driven photocatalytic activity. Appl. Catal. B Environ. 2013, 142, 1–7. [Google Scholar] [CrossRef]
  12. Kumar, O.P.; Shahzad, K.; Nazir, M.A.; Farooq, N.; Malik, M.; Shah, S.S.A.; Ur Rehman, A. Photo-Fenton activated C3N4x/AgOy@ Co1-xBi0. 1-yO7 dual s-scheme heterojunction towards degradation of organic pollutants. Opt. Mater. 2022, 126, 112199. [Google Scholar] [CrossRef]
  13. Kumar, O.P.; Ahmad, M.; Nazir, M.A.; Anum, A.; Jamshaid, M.; Shah, S.S.A.; Rehman, A. Strategic combination of metal–organic frameworks and C3N4 for expeditious photocatalytic degradation of dye pollutants. Environ. Sci. Pollut. Res. 2022, 29, 35300–35313. [Google Scholar] [CrossRef] [PubMed]
  14. Farooq, N.; Luque, R.; Hessien, M.M.; Qureshi, A.M.; Sahiba, F.; Nazir, M.A.; Ur Rehman, A. A comparative study of cerium-and ytterbium-based GO/g-C3N4/Fe2O3 composites for electrochemical and photocatalytic applications. Appl. Sci. 2021, 11, 9000. [Google Scholar] [CrossRef]
  15. Wang, X.; Chen, X.; Thomas, A.; Fu, X.; Antonietti, M. Metal-containing carbon nitride compounds: A new functional organic–metal hybrid material. Adv. Mater. 2009, 21, 1609–1612. [Google Scholar] [CrossRef]
  16. Ye, S.; Wang, R.; Wu, M.; Yuan, Y. A review on g-C3N4 for photocatalytic water splitting and CO2 reduction. Appl. Surf. Sci. 2015, 358, 15–27. [Google Scholar] [CrossRef]
  17. Wang, J.; Wang, S. A critical review on graphitic carbon nitride (g-C3N4)-based materials: Preparation, modification and environmental application. Coord. Chem. Rev. 2022, 453, 214338. [Google Scholar] [CrossRef]
  18. Liebig, J.v. About some nitrogen compounds. Ann. Pharm. 1834, 10, 10. [Google Scholar]
  19. Zhu, J.; Xiao, P.; Li, H.; Carabineiro, S.A. Graphitic carbon nitride: Synthesis, properties, and applications in catalysis. ACS Appl. Mater. Interfaces 2014, 6, 16449–16465. [Google Scholar] [CrossRef]
  20. Franklin, E.C. The ammono carbonic acids. J. Am. Chem. Soc. 1922, 44, 486–509. [Google Scholar] [CrossRef] [Green Version]
  21. Pauling, L.; Sturdivant, J. The structure of cyameluric acid, hydromelonic acid and related substances. Proc. Natl. Acad. Sci. USA 1937, 23, 615–620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Liu, A.Y.; Cohen, M.L. Prediction of new low compressibility solids. Science 1989, 245, 841–842. [Google Scholar] [CrossRef] [PubMed]
  23. David, M.T.; Russell, J.H. Low-compressibility carbon nitrides. Science 1996, 271, 53–55. [Google Scholar]
  24. Maya, L.; Cole, D.R.; Hagaman, E.W. Carbon–nitrogen pyrolyzates: Attempted preparation of carbon nitride. J. Am. Ceram. Soc. 1991, 74, 1686–1688. [Google Scholar] [CrossRef]
  25. Wang, S.; Zhang, J.; Li, B.; Sun, H.; Wang, S. Engineered graphitic carbon nitride-based photocatalysts for visible-light-driven water splitting: A review. Energy Fuels 2021, 35, 6504–6526. [Google Scholar] [CrossRef]
  26. Zhang, Q.; Xiang, X.; Ge, Y.; Yang, C.; Zhang, B.; Deng, K. Selectivity enhancement in the g-C3N4-catalyzed conversion of glucose to gluconic acid and glucaric acid by modification of cobalt thioporphyrazine. J. Catal. 2020, 388, 11–19. [Google Scholar] [CrossRef]
  27. Su, F.; Mathew, S.C.; Lipner, G.; Fu, X.; Antonietti, M.; Blechert, S.; Wang, X. mpg-C3N4-catalyzed selective oxidation of alcohols using O2 and visible light. J. Am. Chem. Soc. 2010, 132, 16299–16301. [Google Scholar] [CrossRef]
  28. Balakrishnan, A.; Chinthala, M. Comprehensive review on advanced reusability of g-C3N4 based photocatalysts for the removal of organic pollutants. Chemosphere 2022, 297, 134190. [Google Scholar] [CrossRef]
  29. Dong, G.; Zhang, Y.; Pan, Q.; Qiu, J. A fantastic graphitic carbon nitride (g-C3N4) material: Electronic structure, photocatalytic and photoelectronic properties. J. Photochem. Photobiol. C Photochem. Rev. 2014, 20, 33–50. [Google Scholar] [CrossRef]
  30. Yan, S.; Li, Z.; Zou, Z. Photodegradation performance of g-C3N4 fabricated by directly heating melamine. Langmuir 2009, 25, 10397–10401. [Google Scholar] [CrossRef]
  31. Liu, J.; Zhang, T.; Wang, Z.; Dawson, G.; Chen, W. Simple pyrolysis of urea into graphitic carbon nitride with recyclable adsorption and photocatalytic activity. J. Mater. Chem. 2011, 21, 14398–14401. [Google Scholar] [CrossRef]
  32. Wu, X.; Chen, F.; Wang, X.; Yu, H. In situ one-step hydrothermal synthesis of oxygen-containing groups-modified g-C3N4 for the improved photocatalytic H2-evolution performance. Appl. Surf. Sci. 2018, 427, 645–653. [Google Scholar] [CrossRef]
  33. Ahmad, I.; Danish, M.; Khan, A.; Muneer, M. One-pot hydrothermal synthesis of a double Z-scheme g-C3N4/AgI/β-AgVO3 ternary nanocomposite for efficient degradation of organic pollutants and DPC–Cr (VI) complex under visible-light irradiation. Photochem. Photobiol. Sci. 2022, 21, 1371–1386. [Google Scholar] [CrossRef] [PubMed]
  34. Sun, J.; Zhang, J.; Zhang, M.; Antonietti, M.; Fu, X.; Wang, X. Bioinspired hollow semiconductor nanospheres as photosynthetic nanoparticles. Nat. Commun. 2012, 3, 1139. [Google Scholar] [CrossRef] [Green Version]
  35. Zhao, S.; Zhang, Y.; Zhou, Y.; Wang, Y.; Qiu, K.; Zhang, C.; Fang, J.; Sheng, X. Facile one-step synthesis of hollow mesoporous g-C3N4 spheres with ultrathin nanosheets for photoredox water splitting. Carbon 2018, 126, 247–256. [Google Scholar] [CrossRef]
  36. Bai, G.; Song, Z.; Geng, H.; Gao, D.; Liu, K.; Wu, S.; Rao, W.; Guo, L.; Wang, J. Oxidized quasi-carbon nitride quantum dots inhibit ice growth. Adv. Mater. 2017, 29, 1606843. [Google Scholar] [CrossRef]
  37. Chen, X.; Liu, Q.; Wu, Q.; Du, P.; Zhu, J.; Dai, S.; Yang, S. Incorporating graphitic carbon nitride (g-C3N4) quantum dots into bulk-heterojunction polymer solar cells leads to efficiency enhancement. Adv. Funct. Mater. 2016, 26, 1719–1728. [Google Scholar] [CrossRef]
  38. Liu, Q.; Zhu, D.; Guo, M.; Yu, Y.; Cao, Y. Facile and efficient fabrication of g-C3N4 quantum dots for fluorescent analysis of trace copper (II) in environmental samples. Chin. Chem. Lett. 2019, 30, 1639–1642. [Google Scholar] [CrossRef]
  39. Yousefzadeh, S.; Morovati, N. Modification of the ultrasonication derived-g-C3N4 nanosheets/quantum dots by MoS2 nanostructures to improve electrocatalytic hydrogen evolution reaction. Int. J. Hydrogen Energy 2020, 45, 33512–33520. [Google Scholar] [CrossRef]
  40. Li, C.; Sun, T.; Zhang, D.; Zhang, X.; Zhang, Y.; Lin, X.; Liu, J.; Zhu, L.; Wang, X.; Shi, Z. Microwave assisted-polymerization synthesis of 0D/2D CNQDs/g-C3N4 with the excellent visible light photocatalytic performance. Mater. Lett. 2022, 308, 131232. [Google Scholar] [CrossRef]
  41. Wang, W.; Jimmy, C.Y.; Shen, Z.; Chan, D.K.; Gu, T. g-C3N4 quantum dots: Direct synthesis, upconversion properties and photocatalytic application. Chem. Commun. 2014, 50, 10148–10150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Dayal, S.; Burda, C. Surface effects on quantum dot-based energy transfer. J. Am. Chem. Soc. 2007, 129, 7977–7981. [Google Scholar] [CrossRef] [PubMed]
  43. Chen, Q.; Tong, R.; Chen, X.; Xue, Y.; Xie, Z.; Kuang, Q.; Zheng, L. Ultrafine ZnO quantum dot-modified TiO2 composite photocatalysts: The role of the quantum size effect in heterojunction-enhanced photocatalytic hydrogen evolution. Catal. Sci. Technol. 2018, 8, 1296–1303. [Google Scholar] [CrossRef]
  44. Yeo, I.; Malik, N.S.; Munsch, M.; Dupuy, E.; Bleuse, J.; Niquet, Y.-M.; Gérard, J.-M.; Claudon, J.; Wagner, É.; Seidelin, S. Surface effects in a semiconductor photonic nanowire and spectral stability of an embedded single quantum dot. Appl. Phys. Lett. 2011, 99, 233106. [Google Scholar] [CrossRef] [Green Version]
  45. Cui, Y.; Ding, Z.; Fu, X.; Wang, X. Construction of conjugated carbon nitride nanoarchitectures in solution at low temperatures for photoredox catalysis. Angew. Chem. Int. Ed. 2012, 51, 11814–11818. [Google Scholar] [CrossRef]
  46. Wang, Q.; Shi, Y.; Du, Z.; He, J.; Zhong, J.; Zhao, L.; She, H.; Liu, G.; Su, B. Synthesis of rod-like g-C3N4/ZnS composites with superior photocatalytic activity for the degradation of methyl orange. Eur. J. Inorg. Chem. 2015, 2015, 4108–4115. [Google Scholar] [CrossRef]
  47. Bai, X.; Wang, L.; Zong, R.; Zhu, Y. Photocatalytic activity enhanced via g-C3N4 nanoplates to nanorods. J. Phys. Chem. C 2013, 117, 9952–9961. [Google Scholar] [CrossRef]
  48. Mo, Z.; Xu, H.; Chen, Z.; She, X.; Song, Y.; Wu, J.; Yan, P.; Xu, L.; Lei, Y.; Yuan, S. Self-assembled synthesis of defect-engineered graphitic carbon nitride nanotubes for efficient conversion of solar energy. Appl. Catal. B Environ. 2018, 225, 154–161. [Google Scholar] [CrossRef]
  49. Jiang, Z.; Zhang, X.; Chen, H.S.; Hu, X.; Yang, P. Formation of g-C3N4 nanotubes towards superior photocatalysis performance. ChemCatChem 2019, 11, 4558–4567. [Google Scholar] [CrossRef]
  50. Chen, D.; Wang, K.; Xiang, D.; Zong, R.; Yao, W.; Zhu, Y. Significantly enhancement of photocatalytic performances via core–shell structure of ZnO@ mpg-C3N4. Appl. Catal. B Environ. 2014, 147, 554–561. [Google Scholar] [CrossRef]
  51. Zhou, Y.; Wang, Z.; Huang, L.; Zaman, S.; Lei, K.; Yue, T.; Li, Z.a.; You, B.; Xia, B.Y. Engineering 2D photocatalysts toward carbon dioxide reduction. Adv. Energy Mater. 2021, 11, 2003159. [Google Scholar] [CrossRef]
  52. Niu, P.; Zhang, L.; Liu, G.; Cheng, H.M. Graphene-like carbon nitride nanosheets for improved photocatalytic activities. Adv. Funct. Mater. 2012, 22, 4763–4770. [Google Scholar] [CrossRef]
  53. Liu, Q.; Wang, X.; Yang, Q.; Zhang, Z.; Fang, X. Mesoporous g-C3N4 nanosheets prepared by calcining a novel supramolecular precursor for high-efficiency photocatalytic hydrogen evolution. Appl. Surf. Sci. 2018, 450, 46–56. [Google Scholar] [CrossRef]
  54. Qian, X.; Meng, X.; Sun, J.; Jiang, L.; Wang, Y.; Zhang, J.; Hu, X.; Shalom, M.; Zhu, J. Salt-assisted synthesis of 3D porous g-C3N4 as a bifunctional photo-and electrocatalyst. ACS Appl. Mater. Interfaces 2019, 11, 27226–27232. [Google Scholar] [CrossRef] [PubMed]
  55. Toe, C.Y.; Tsounis, C.; Zhang, J.; Masood, H.; Gunawan, D.; Scott, J.; Amal, R. Advancing photoreforming of organics: Highlights on photocatalyst and system designs for selective oxidation reactions. Energy Environ. Sci. 2021, 14, 1140–1175. [Google Scholar] [CrossRef]
  56. Gao, Y.; Zhao, W.; Zhang, D. Preparation of prismatic g-C3N4 by CTAB hydrothermal method and its degradation performance under visible light. Diam. Relat. Mater. 2022, 121, 108787. [Google Scholar] [CrossRef]
  57. Lin, B.; Xia, M.; Xu, B.; Chong, B.; Chen, Z.; Yang, G. Bio-inspired nanostructured g-C3N4-based photocatalysts: A comprehensive review. Chin. J. Catal. 2022, 43, 2141–2172. [Google Scholar] [CrossRef]
  58. Zhao, Z.; Wang, X.; Wang, S.; Xiao, Z.; Zhai, S.; Ma, J.; Dong, X.; Sun, H.; An, Q. Three-dimensional hierarchical seaweed-derived carbonaceous network with designed g-C3N4 nanosheets: Preparation and mechanism insight for 4-nitrophenol photoreduction. Langmuir 2022, 38, 11054–11067. [Google Scholar] [CrossRef]
  59. Liu, Y.; Guo, X.; Chen, Z.; Zhang, W.; Wang, Y.; Zheng, Y.; Tang, X.; Zhang, M.; Peng, Z.; Li, R. Microwave-synthesis of g-C3N4 nanoribbons assembled seaweed-like architecture with enhanced photocatalytic property. Appl. Catal. B Environ. 2020, 266, 118624. [Google Scholar] [CrossRef]
  60. Jiang, H.; Li, Y.; Wang, D.; Hong, X.; Liang, B. Recent advances in heteroatom doped graphitic carbon nitride (g-C3N4) and g-C3N4/metal oxide composite photocatalysts. Curr. Org. Chem. 2020, 24, 673–693. [Google Scholar] [CrossRef]
  61. Liu, X.; Ma, R.; Zhuang, L.; Hu, B.; Chen, J.; Liu, X.; Wang, X. Recent developments of doped g-C3N4 photocatalysts for the degradation of organic pollutants. Crit. Rev. Environ. Sci. Technol. 2021, 51, 751–790. [Google Scholar] [CrossRef]
  62. He, F.; Wang, Z.; Li, Y.; Peng, S.; Liu, B. The nonmetal modulation of composition and morphology of g-C3N4-based photocatalysts. Appl. Catal. B Environ. 2020, 269, 118828. [Google Scholar] [CrossRef]
  63. Huang, Z.; Song, J.; Pan, L.; Wang, Z.; Zhang, X.; Zou, J.-J.; Mi, W.; Zhang, X.; Wang, L. Carbon nitride with simultaneous porous network and O-doping for efficient solar-energy-driven hydrogen evolution. Nano Energy 2015, 12, 646–656. [Google Scholar] [CrossRef]
  64. Zhang, J.-W.; Gong, S.; Mahmood, N.; Pan, L.; Zhang, X.; Zou, J.-J. Oxygen-doped nanoporous carbon nitride via water-based homogeneous supramolecular assembly for photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2018, 221, 9–16. [Google Scholar] [CrossRef]
  65. She, X.; Liu, L.; Ji, H.; Mo, Z.; Li, Y.; Huang, L.; Du, D.; Xu, H.; Li, H. Template-free synthesis of 2D porous ultrathin nonmetal-doped g-C3N4 nanosheets with highly efficient photocatalytic H2 evolution from water under visible light. Appl. Catal. B Environ. 2016, 187, 144–153. [Google Scholar] [CrossRef]
  66. Ma, X.; Lv, Y.; Xu, J.; Liu, Y.; Zhang, R.; Zhu, Y. A strategy of enhancing the photoactivity of g-C3N4 via doping of nonmetal elements: A first-principles study. J. Phys. Chem. C 2012, 116, 23485–23493. [Google Scholar] [CrossRef]
  67. Wang, Y.; Zhao, S.; Zhang, Y.; Fang, J.; Zhou, Y.; Yuan, S.; Zhang, C.; Chen, W. One-pot synthesis of K-doped g-C3N4 nanosheets with enhanced photocatalytic hydrogen production under visible-light irradiation. Appl. Surf. Sci. 2018, 440, 258–265. [Google Scholar] [CrossRef]
  68. Gao, J.; Wang, Y.; Zhou, S.; Lin, W.; Kong, Y. A facile one-step synthesis of Fe-doped g-C3N4 nanosheets and their improved visible-light photocatalytic performance. ChemCatChem 2017, 9, 1708–1715. [Google Scholar] [CrossRef] [Green Version]
  69. Deng, P.; Shi, L.; Wang, H.; Qi, W. One-step preparation of novel K+ and cyano-group co-doped crystalline polymeric carbon nitride with highly efficient H2 evolution. Colloids Surf. A Physicochem. Eng. Asp. 2020, 601, 125023. [Google Scholar] [CrossRef]
  70. Xu, L.; Zeng, J.; Li, Q.; Xia, L.; Luo, X.; Ma, Z.; Peng, B.; Xiong, S.; Li, Z.; Wang, L.-L. Defect-engineered 2D/2D hBN/g-C3N4 Z-scheme heterojunctions with full visible-light absorption: Efficient metal-free photocatalysts for hydrogen evolution. Appl. Surf. Sci. 2021, 547, 149207. [Google Scholar] [CrossRef]
  71. Wang, M.; Jin, C.; Kang, J.; Liu, J.; Tang, Y.; Li, Z.; Li, S. CuO/g-C3N4 2D/2D heterojunction photocatalysts as efficient peroxymonosulfate activators under visible light for oxytetracycline degradation: Characterization, efficiency and mechanism. Chem. Eng. J. 2021, 416, 128118. [Google Scholar] [CrossRef]
  72. Zhao, W.; Li, Y.; Zhao, P.; Zhang, L.; Dai, B.; Huang, H.; Zhou, J.; Zhu, Y.; Ma, K.; Leung, D.Y. Insights into the photocatalysis mechanism of the novel 2D/3D Z-Scheme g-C3N4/SnS2 heterojunction photocatalysts with excellent photocatalytic performances. J. Hazard. Mater. 2021, 402, 123711. [Google Scholar] [CrossRef] [PubMed]
  73. Wang, K.; Jiang, L.; Wu, X.; Zhang, G. Vacancy mediated Z-scheme charge transfer in a 2D/2D La2Ti2O7/g-C3N4 nanojunction as a bifunctional photocatalyst for solar-to-energy conversion. J. Mater. Chem. A 2020, 8, 13241–13247. [Google Scholar] [CrossRef]
  74. Chen, X.; Wang, J.; Chai, Y.; Zhang, Z.; Zhu, Y. Efficient photocatalytic overall water splitting induced by the giant internal electric field of ag-C3N4/rGO/PDIP Z-scheme heterojunction. Adv. Mater. 2021, 33, 2007479. [Google Scholar] [CrossRef] [PubMed]
  75. Xu, H.; Wu, L.; Jin, L.; Wu, K. Combination mechanism and enhanced visible-light photocatalytic activity and stability of CdS/g-C3N4 heterojunctions. J. Mater. Sci. Technol. 2017, 33, 30–38. [Google Scholar] [CrossRef]
  76. Liu, Y.; Li, X.; He, H.; Yang, S.; Jia, G.; Liu, S. CoP imbedded g-C3N4 heterojunctions for highly efficient photo, electro and photoelectrochemical water splitting. J. Colloid Interface Sci. 2021, 599, 23–33. [Google Scholar] [CrossRef] [PubMed]
  77. Ma, J.; Zhou, W.; Tan, X.; Yu, T. Potassium ions intercalated into g-C3N4-modified TiO2 nanobelts for the enhancement of photocatalytic hydrogen evolution activity under visible-light irradiation. Nanotechnology 2018, 29, 215706. [Google Scholar] [CrossRef]
  78. Monsef, R.; Ghiyasiyan-Arani, M.; Salavati-Niasari, M. Design of magnetically recyclable ternary Fe2O3/EuVO4/g-C3N4 nanocomposites for photocatalytic and electrochemical hydrogen storage. ACS Appl. Energy Mater. 2021, 4, 680–695. [Google Scholar] [CrossRef]
  79. Guo, F.; Huang, X.; Chen, Z.; Cao, L.; Cheng, X.; Chen, L.; Shi, W. Construction of Cu3P-ZnSnO3-g-C3N4 pnn heterojunction with multiple built-in electric fields for effectively boosting visible-light photocatalytic degradation of broad-spectrum antibiotics. Sep. Purif. Technol. 2021, 265, 118477. [Google Scholar] [CrossRef]
  80. Li, K.; Chen, J.; Ao, Y.; Wang, P. Preparation of a ternary g-C3N4-CdS/Bi4O5I2 composite photocatalysts with two charge transfer pathways for efficient degradation of acetaminophen under visible light irradiation. Sep. Purif. Technol. 2021, 259, 118177. [Google Scholar] [CrossRef]
  81. Xiang, W.; Ji, Q.; Xu, C.; Guo, Y.; Liu, Y.; Sun, D.; Zhou, W.; Xu, Z.; Qi, C.; Yang, S. Accelerated photocatalytic degradation of iohexol over Co3O4/g-C3N4/Bi2O2CO3 of pn/nn dual heterojunction under simulated sunlight by persulfate. Appl. Catal. B Environ. 2021, 285, 119847. [Google Scholar] [CrossRef]
  82. Liu, S.; Li, F.; Li, Y.; Hao, Y.; Wang, X.; Li, B.; Liu, R. Fabrication of ternary g-C3N4/Al2O3/ZnO heterojunctions based on cascade electron transfer toward molecular oxygen activation. Appl. Catal. B Environ. 2017, 212, 115–128. [Google Scholar] [CrossRef]
  83. Liu, Y.; Tian, J.; Wei, L.; Wang, Q.; Wang, C.; Xing, Z.; Li, X.; Yang, W.; Yang, C. Modified g-C3N4/TiO2/CdS ternary heterojunction nanocomposite as highly visible light active photocatalyst originated from CdS as the electron source of TiO2 to accelerate Z-type heterojunction. Sep. Purif. Technol. 2021, 257, 117976. [Google Scholar] [CrossRef]
  84. Li, Z.; Huang, F.; Xu, Y.; Yan, A.; Dong, H.; Luo, S.; Hu, M. 2D/2D Nb3O7F/g-C3N4 heterojunction photocatalysts with enhanced hydrogen evolution activity. ACS Appl. Energy Mater. 2021, 4, 839–845. [Google Scholar] [CrossRef]
  85. Shi, J.; Mao, L.; Cai, C.; Li, G.; Cheng, C.; Zheng, B.; Hu, Y.; Huang, Z.; Hu, X.; Żyła, G. One-pot fabrication of 2D/2D HCa2Nb3O10/g-C3N4 type II heterojunctions towards enhanced photocatalytic H2 evolution under visible-light irradiation. Catal. Sci. Technol. 2020, 10, 5896–5902. [Google Scholar] [CrossRef]
  86. Bai, L.; Huang, H.; Zhang, S.; Hao, L.; Zhang, Z.; Li, H.; Sun, L.; Guo, L.; Huang, H.; Zhang, Y. Photocatalysis-assisted Co3O4/g-C3N4 p–n junction all-solid-state supercapacitors: A bridge between energy storage and photocatalysis. Adv. Sci. 2020, 7, 2001939. [Google Scholar] [CrossRef]
  87. Vesali-Kermani, E.; Habibi-Yangjeh, A.; Diarmand-Khalilabad, H.; Ghosh, S. Nitrogen photofixation ability of g-C3N4 nanosheets/Bi2MoO6 heterojunction photocatalyst under visible-light illumination. J. Colloid Interface Sci. 2020, 563, 81–91. [Google Scholar] [CrossRef]
  88. Kumaresan, N.; Sinthiya, M.M.A.; Sarathbavan, M.; Ramamurthi, K.; Sethuraman, K.; Babu, R.R. Synergetic effect of g-C3N4/ZnO binary nanocomposites heterojunction on improving charge carrier separation through 2D/1D nanostructures for effective photocatalytic activity under the sunlight irradiation. Sep. Purif. Technol. 2020, 244, 116356. [Google Scholar] [CrossRef]
  89. Yan, T.; Liu, H.; Jin, Z. g-C3N4/α-Fe2O3 supported zero-dimensional Co3S4 nanoparticles form S-scheme heterojunction photocatalyst for efficient hydrogen production. Energy Fuels 2020, 35, 856–867. [Google Scholar] [CrossRef]
  90. Fu, Y.; Zhang, Y.; Xie, X.; Wang, H.; Wei, L.; Ma, M.; Yan, Q. Functionalized carbon nanotube bridge interface drove Bi2O2CO3/g-C3N4 S-scheme heterojunction with enhanced visible-light photocatalytic activity. Sep. Purif. Technol. 2021, 274, 119032. [Google Scholar] [CrossRef]
  91. Guo, H.; Niu, C.-G.; Yang, Y.-Y.; Liang, C.; Niu, H.-Y.; Liu, H.-Y.; Li, L.; Tang, N. Interfacial Co-N bond bridged CoB/g-C3N4 Schottky junction with modulated charge transfer dynamics for highly efficient photocatalytic Staphylococcus aureus inactivation. Chem. Eng. J. 2021, 422, 130029. [Google Scholar] [CrossRef]
  92. Lin, S.; Zhang, N.; Wang, F.; Lei, J.; Zhou, L.; Liu, Y.; Zhang, J. Carbon vacancy mediated incorporation of Ti3C2 quantum dots in a 3D inverse opal g-C3N4 Schottky junction catalyst for photocatalytic H2O2 production. ACS Sustain. Chem. Eng. 2020, 9, 481–488. [Google Scholar] [CrossRef]
  93. Yi, X.; Yuan, J.; Tang, H.; Du, Y.; Hassan, B.; Yin, K.; Chen, Y.; Liu, X. Embedding few-layer Ti3C2Tx into alkalized g-C3N4 nanosheets for efficient photocatalytic degradation. J. Colloid Interface Sci. 2020, 571, 297–306. [Google Scholar] [CrossRef] [PubMed]
  94. Ye, J.; Liu, J.; An, Y. Electric field and strain effects on the electronic and optical properties of g-C3N4/WSe2 van der Waals heterostructure. Appl. Surf. Sci. 2020, 501, 144262. [Google Scholar] [CrossRef]
  95. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-scheme heterojunction photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  96. Fu, J.; Xu, Q.; Low, J.; Jiang, C.; Yu, J. Ultrathin 2D/2D WO3/g-C3N4 step-scheme H2-production photocatalyst. Appl. Catal. B Environ. 2019, 243, 556–565. [Google Scholar] [CrossRef]
  97. Ren, Y.; Li, Y.; Wu, X.; Wang, J.; Zhang, G. S-scheme Sb2WO6/g-C3N4 photocatalysts with enhanced visible-light-induced photocatalytic NO oxidation performance. Chin. J. Catal. 2021, 42, 69–77. [Google Scholar] [CrossRef]
  98. Wang, J.; Wang, G.; Cheng, B.; Yu, J.; Fan, J. Sulfur-doped g-C3N4/TiO2 S-scheme heterojunction photocatalyst for Congo Red photodegradation. Chin. J. Catal. 2021, 42, 56–68. [Google Scholar] [CrossRef]
  99. Dai, Z.; Zhen, Y.; Sun, Y.; Li, L.; Ding, D. ZnFe2O4/g-C3N4 S-scheme photocatalyst with enhanced adsorption and photocatalytic activity for uranium (VI) removal. Chem. Eng. J. 2021, 415, 129002. [Google Scholar]
  100. Mei, F.; Li, Z.; Dai, K.; Zhang, J.; Liang, C. Step-scheme porous g-C3N4/Zn0.2Cd0.8S-DETA composites for efficient and stable photocatalytic H2 production. Chin. J. Catal. 2020, 41, 41–49. [Google Scholar] [CrossRef]
  101. Xie, Q.; He, W.; Liu, S.; Li, C.; Zhang, J.; Wong, P.K. Bifunctional S-scheme g-C3N4/Bi/BiVO4 hybrid photocatalysts toward artificial carbon cycling. Chin. J. Catal. 2020, 41, 140–153. [Google Scholar] [CrossRef]
  102. Du, X.; Song, S.; Wang, Y.; Jin, W.; Ding, T.; Tian, Y.; Li, X. Facile one-pot synthesis of defect-engineered step-scheme WO3/g-C3N4 heterojunctions for efficient photocatalytic hydrogen production. Catal. Sci. Technol. 2021, 11, 2734–2744. [Google Scholar] [CrossRef]
  103. Wang, X.-J.; Tian, X.; Sun, Y.-J.; Zhu, J.-Y.; Li, F.-T.; Mu, H.-Y.; Zhao, J. Enhanced Schottky effect of a 2D–2D CoP/g-C3N4 interface for boosting photocatalytic H2 evolution. Nanoscale 2018, 10, 12315–12321. [Google Scholar] [CrossRef]
  104. She, X.; Xu, H.; Li, L.; Mo, Z.; Zhu, X.; Yu, Y.; Song, Y.; Wu, J.; Qian, J.; Yuan, S. Steering charge transfer for boosting photocatalytic H2 evolution: Integration of two-dimensional semiconductor superiorities and noble-metal-free Schottky junction effect. Appl. Catal. B Environ. 2019, 245, 477–485. [Google Scholar] [CrossRef]
  105. Yang, Y.; Zeng, Z.; Zeng, G.; Huang, D.; Xiao, R.; Zhang, C.; Zhou, C.; Xiong, W.; Wang, W.; Cheng, M. Ti3C2 Mxene/porous g-C3N4 interfacial Schottky junction for boosting spatial charge separation in photocatalytic H2O2 production. Appl. Catal. B Environ. 2019, 258, 117956. [Google Scholar] [CrossRef]
  106. Shaheer, A.M.; Karthik, P.; Karthik, G.; Shankar, M.; Neppolian, B. Dual role of a g-C3N4/carbon intra-Schottky junction in charge carrier generation and separation for efficient solar H2 production. Catal. Sci. Technol. 2019, 9, 3493–3503. [Google Scholar] [CrossRef]
  107. Li, H.J.W.; Zhou, H.; Chen, K.; Liu, K.; Li, S.; Jiang, K.; Zhang, W.; Xie, Y.; Cao, Z.; Li, H. Metallic MoO2-modified graphitic carbon nitride boosting photocatalytic CO2 reduction via Schottky junction. Solar Rrl 2020, 4, 1900416. [Google Scholar] [CrossRef]
  108. Sun, Z.; Fang, W.; Zhao, L.; Wang, H. 3D porous Cu-NPs/g-C3N4 foam with excellent CO2 adsorption and Schottky junction effect for photocatalytic CO2 reduction. Appl. Surf. Sci. 2020, 504, 144347. [Google Scholar] [CrossRef]
  109. Li, J.; Zhang, Z.; Cui, W.; Wang, H.; Cen, W.; Johnson, G.; Jiang, G.; Zhang, S.; Dong, F. The spatially oriented charge flow and photocatalysis mechanism on internal van der Waals heterostructures enhanced g-C3N4. ACS Catal. 2018, 8, 8376–8385. [Google Scholar] [CrossRef]
  110. Ran, J.; Guo, W.; Wang, H.; Zhu, B.; Yu, J.; Qiao, S.Z. Metal-free 2D/2D phosphorene/g-C3N4 Van der Waals heterojunction for highly enhanced visible-light photocatalytic H2 production. Adv. Mater. 2018, 30, 1800128. [Google Scholar]
  111. Sun, D.; Chi, D.; Yang, Z.; Xing, Z.; Yin, J.; Li, Z.; Zhu, Q.; Zhou, W. Mesoporous g-C3N4/Zn–Ti LDH laminated van der Waals heterojunction nanosheets as remarkable visible-light-driven photocatalysts. Int. J. Hydrogen Energy 2019, 44, 16348–16358. [Google Scholar] [CrossRef]
  112. Dong, H.; Hong, S.; Zhang, P.; Yu, S.; Wang, Y.; Yuan, S.; Li, H.; Sun, J.; Chen, G.; Li, C. Metal-free Z-scheme 2D/2D VdW heterojunction for high-efficiency and durable photocatalytic H2 production. Chem. Eng. J. 2020, 395, 125150. [Google Scholar] [CrossRef]
  113. Song, X.; Wu, Y.; Zhang, X.; Li, X.; Zhu, Z.; Ma, C.; Yan, Y.; Huo, P.; Yang, G. Boosting charge carriers separation and migration efficiency via fabricating all organic van der Waals heterojunction for efficient photoreduction of CO2. Chem. Eng. J. 2021, 408, 127292. [Google Scholar] [CrossRef]
  114. Han, Z.; Fu, Y.; Zhang, Y.; Zhang, X.; Meng, X.; Zhou, Z.; Su, Z. Metal–organic framework (MOF) composite materials for photocatalytic CO2 reduction under visible light. Dalton Trans. 2021, 50, 3186–3192. [Google Scholar] [CrossRef] [PubMed]
  115. Zhang, X.; Yu, X.; Li, J.; Xu, L.; Li, B. Construction of MOFs/g-C3N4 composite for accelerating visible-light-driven hydrogen evolution. Int. J. Hydrogen Energy 2022, 47, 18007–18017. [Google Scholar] [CrossRef]
  116. Lv, P.; Duan, F.; Sheng, J.; Lu, S.; Zhu, H.; Du, M.; Chen, M. The 2D/2D p–n heterojunction of ZnCoMOF/g-C3N4 with enhanced photocatalytic hydrogen evolution under visible light irradiation. Appl. Organomet. Chem. 2021, 35, e6124. [Google Scholar] [CrossRef]
  117. Zhang, Y.; Zhou, J.; Feng, Q.; Chen, X.; Hu, Z. Visible light photocatalytic degradation of MB using UiO-66/g-C3N4 heterojunction nanocatalyst. Chemosphere 2018, 212, 523–532. [Google Scholar] [CrossRef]
  118. Ma, X.; Li, X.; Li, M.; Ma, X.; Yu, L.; Dai, Y. Effect of the structure distortion on the high photocatalytic performance of C60/g-C3N4 composite. Appl. Surf. Sci. 2017, 414, 124–130. [Google Scholar] [CrossRef]
  119. Zhang, Y.; Wang, G.; Jin, Z. An orderly assembled g-C3N4, rGO and Ni2P photocatalyst for efficient hydrogen evolution. Int. J. Hydrogen Energy 2019, 44, 10316–10327. [Google Scholar] [CrossRef]
  120. Liu, J.; Song, Y.; Xu, H.; Zhu, X.; Lian, J.; Xu, Y.; Zhao, Y.; Huang, L.; Ji, H.; Li, H. Non-metal photocatalyst nitrogen-doped carbon nanotubes modified mpg-C3N4: Facile synthesis and the enhanced visible-light photocatalytic activity. J. Colloid Interface Sci. 2017, 494, 38–46. [Google Scholar] [CrossRef]
  121. Cheng, L.; Zhang, H.; Li, X.; Fan, J.; Xiang, Q. Carbon–graphitic carbon nitride hybrids for heterogeneous photocatalysis. Small 2021, 17, 2005231. [Google Scholar] [CrossRef] [PubMed]
  122. Yuan, B.; Wei, J.; Hu, T.; Yao, H.; Jiang, Z.; Fang, Z.; Chu, Z. Simple synthesis of g-C3N4/rGO hybrid catalyst for the photocatalytic degradation of rhodamine B. Chin. J. Catal. 2015, 36, 1009–1016. [Google Scholar] [CrossRef]
  123. Ge, L.; Han, C. Synthesis of MWNTs/g-C3N4 composite photocatalysts with efficient visible light photocatalytic hydrogen evolution activity. Appl. Catal. B Environ. 2012, 117, 268–274. [Google Scholar] [CrossRef]
  124. Wang, Q.; Feng, Z.; He, H.; Hu, X.; Mao, J.; Chen, X.; Liu, L.; Wei, X.; Liu, D.; Bi, S. Nonblinking carbon dots for imaging and tracking receptors on a live cell membrane. Chem. Commun. 2021, 57, 5554–5557. [Google Scholar]
  125. Zhang, F.; Wen, Q.; Hong, M.; Zhuang, Z.; Yu, Y. Efficient and sustainable metal-free GR/C3N4/CDots ternary heterostructrues for versatile visible-light-driven photoredox applications: Toward synergistic interaction of carbon materials. Chem. Eng. J. 2017, 307, 593–603. [Google Scholar]
  126. He, H.; Liu, L.; Chen, X.; Wang, Q.; Wang, X.; Nau, W.M.; Huang, F. Carbon dot blinking enables accurate molecular counting at nanoscale resolution. Anal. Chem. 2021, 93, 3968–3975. [Google Scholar] [CrossRef]
  127. Liu, J.; Liu, Y.; Liu, N.; Han, Y.; Zhang, X.; Huang, H.; Lifshitz, Y.; Lee, S.-T.; Zhong, J.; Kang, Z. Metal-free efficient photocatalyst for stable visible water splitting via a two-electron pathway. Science 2015, 347, 970–974. [Google Scholar] [CrossRef]
  128. Asadzadeh-Khaneghah, S.; Habibi-Yangjeh, A. g-C3N4/carbon dot-based nanocomposites serve as efficacious photocatalysts for environmental purification and energy generation: A review. J. Clean. Prod. 2020, 276, 124319. [Google Scholar] [CrossRef]
  129. Fang, S.; Xia, Y.; Lv, K.; Li, Q.; Sun, J.; Li, M. Effect of carbon-dots modification on the structure and photocatalytic activity of g-C3N4. Appl. Catal. B Environ. 2016, 185, 225–232. [Google Scholar] [CrossRef]
  130. Ai, L.; Shi, R.; Yang, J.; Zhang, K.; Zhang, T.; Lu, S. Efficient combination of g-C3N4 and CDs for enhanced photocatalytic performance: A review of synthesis, strategies, and applications. Small 2021, 17, 2007523. [Google Scholar] [CrossRef]
  131. Yan, L.; Zhong, S.; Igou, T.; Gao, H.; Li, J.; Chen, Y. Development of machine learning models to enhance element-doped g-C3N4 photocatalyst for hydrogen production through splitting water. Int. J. Hydrogen Energy 2022, 47, 34075–34089. [Google Scholar] [CrossRef]
  132. Sharma, R.; Almáši, M.; Nehra, S.P.; Rao, V.S.; Panchal, P.; Paul, D.R.; Jain, I.P.; Sharma, A. Photocatalytic hydrogen production using graphitic carbon nitride (GCN): A precise review. Renew. Sustain. Energy Rev. 2022, 168, 112776. [Google Scholar] [CrossRef]
  133. Kar, S.; Bishwal, L.; Kommula, B.; Bhattacharyya, S. Critical optimization of phosphorus functionalized carbon nanomaterials for metal-free solar hydrogen production and simultaneous organic transformation. Adv. Opt. Mater. 2022, 10, 2102641. [Google Scholar] [CrossRef]
  134. Gao, Y.; Hou, F.; Hu, S.; Wu, B.; Wang, Y.; Zhang, H.; Jiang, B.; Fu, H. Graphene quantum-dot-modified hexagonal tubular carbon nitride for visible-light photocatalytic hydrogen evolution. ChemCatChem 2018, 10, 1330–1335. [Google Scholar] [CrossRef]
  135. Ling, G.Z.S.; Ng, S.F.; Ong, W.J. Tailor-engineered 2D cocatalysts: Harnessing electron–hole redox center of 2D g-C3N4 photocatalysts toward solar-to-chemical conversion and environmental purification. Adv. Funct. Mater. 2022, 32, 2111875. [Google Scholar] [CrossRef]
  136. Shi, W.; Wang, J.; Yang, S.; Lin, X.; Guo, F.; Shi, J. Fabrication of a ternary carbon dots/CoO/g-C3N4 nanocomposite photocatalyst with enhanced visible-light-driven photocatalytic hydrogen production. J. Chem. Technol. Biotechnol. 2020, 95, 2129–2138. [Google Scholar] [CrossRef]
  137. Zhang, S.; Gao, M.; Zhai, Y.; Wen, J.; Yu, J.; He, T.; Kang, Z.; Lu, S. Which kind of nitrogen chemical states doped carbon dots loaded by g-C3N4 is the best for photocatalytic hydrogen production. J. Colloid Interface Sci. 2022, 622, 662–674. [Google Scholar] [CrossRef]
  138. Hussien, M.K.; Sabbah, A.; Qorbani, M.; Elsayed, M.H.; Raghunath, P.; Lin, T.-Y.; Quadir, S.; Wang, H.-Y.; Wu, H.-L.; Tzou, D.-L.M. Metal-free four-in-one modification of g-C3N4 for superior photocatalytic CO2 reduction and H2 evolution. Chem. Eng. J. 2022, 430, 132853. [Google Scholar] [CrossRef]
  139. Li, L.; Wang, X.; Gu, H.; Zhang, H.; Zhang, J.; Zhang, Q.; Dai, W.-L. Which Is more efficient in promoting the photocatalytic H2 evolution performance of g-C3N4: Monometallic nanocrystal, heterostructural nanocrystal, or bimetallic nanocrystal? Inorg. Chem. 2022, 61, 4760–4768. [Google Scholar] [CrossRef]
  140. Sun, Y.; Kumar, V.; Kim, K.-H. The assessment of graphitic carbon nitride (g-C3N4) materials for hydrogen evolution reaction: Effect of metallic and non-metallic modifications. Sep. Purif. Technol. 2022, 305, 122413. [Google Scholar] [CrossRef]
  141. Devarayapalli, K.; Vattikuti, S.P.; Sreekanth, T.; Yoo, K.S.; Nagajyothi, P.; Shim, J. Hydrogen production and photocatalytic activity of g-C3N4/Co-MOF (ZIF-67) nanocomposite under visible light irradiation. Appl. Organomet. Chem. 2020, 34, e5376. [Google Scholar] [CrossRef]
  142. Yu, H.; Ma, H.; Wu, X.; Wang, X.; Fan, J.; Yu, J. One-step realization of crystallization and cyano-group generation for g-C3N4 photocatalysts with improved H2 production. Solar Rrl 2021, 5, 2000372. [Google Scholar] [CrossRef]
  143. Wang, K.; Li, Y.; Li, J.; Zhang, G. Boosting interfacial charge separation of Ba5Nb4O15/g-C3N4 photocatalysts by 2D/2D nanojunction towards efficient visible-light driven H2 generation. Appl. Catal. B Environ. 2020, 263, 117730. [Google Scholar] [CrossRef]
  144. Katsumata, H.; Tateishi, I.; Furukawa, M.; Kaneco, S. Highly photocatalytic hydrogen generation over P-doped g-C3N4 with aromatic ring structure. Mater. Lett. 2021, 299, 130068. [Google Scholar] [CrossRef]
  145. Xi, C.; Deng, P.; Miao, D.; Zhang, L.; Liu, W.; Hou, Y. Improved charge transfer by binary Ni3S2-NiS2 for boosting the photocatalytic hydrogen generation over g-C3N4. Mater. Lett. 2022, 307, 131012. [Google Scholar] [CrossRef]
  146. Liang, G.; Waqas, M.; Yang, B.; Xiao, K.; Li, J.; Zhu, C.; Zhang, J.; Duan, H. Enhanced photocatalytic hydrogen evolution under visible light irradiation by p-type MoS2/n-type Ni2P doped g-C3N4. Appl. Surf. Sci. 2020, 504, 144448. [Google Scholar] [CrossRef]
  147. Nasir, M.S.; Yang, G.; Ayub, I.; Wang, X.; Wang, S.; Yan, W. Hybridization of g-C3N4 quantum dots with 1D branched TiO2 fiber for efficient visible light-driven photocatalytic hydrogen generation. Int. J. Hydrogen Energy 2020, 45, 13994–14005. [Google Scholar] [CrossRef]
  148. Samsudin, M.F.R.; Ullah, H.; Bashiri, R.; Mohamed, N.M.; Sufian, S.; Ng, Y.H. Experimental and DFT insights on microflower g-C3N4/BiVO4 photocatalyst for enhanced photoelectrochemical hydrogen generation from lake water. ACS Sustain. Chem. Eng. 2020, 8, 9393–9403. [Google Scholar] [CrossRef]
  149. Qi, K.; Lv, W.; Khan, I.; Liu, S.-y. Photocatalytic H2 generation via CoP quantum-dot-modified g-C3N4 synthesized by electroless plating. Chin. J. Catal. 2020, 41, 114–121. [Google Scholar] [CrossRef]
  150. Ma, B.; Zhao, J.; Ge, Z.; Chen, Y.; Yuan, Z. 5 nm NiCoP nanoparticles coupled with g-C3N4 as high-performance photocatalyst for hydrogen evolution. Sci. China Mater. 2020, 63, 258–266. [Google Scholar] [CrossRef] [Green Version]
  151. Jia, J.; Sun, W.; Zhang, Q.; Zhang, X.; Hu, X.; Liu, E.; Fan, J. Inter-plane heterojunctions within 2D/2D FeSe2/g-C3N4 nanosheet semiconductors for photocatalytic hydrogen generation. Appl. Catal. B Environ. 2020, 261, 118249. [Google Scholar] [CrossRef]
  152. Fan, Y.; Chen, H.; Cui, D.; Fan, Z.; Xue, C. NiCo2O4/g-C3N4 Heterojunction Photocatalyst for Efficient H2 Generation. Energy Technol. 2021, 9, 2000973. [Google Scholar] [CrossRef]
  153. Bi, Z.-x.; Guo, R.-t.; Ji, X.-y.; Hu, X.; Wang, J.; Chen, X.; Pan, W.-g. Direct Z-scheme CoS/g-C3N4 heterojunction with NiS co-catalyst for efficient photocatalytic hydrogen generation. Int. J. Hydrogen Energy 2022, 47, 34430–34443. [Google Scholar] [CrossRef]
  154. Zada, A.; Khan, M.; Hussain, Z.; Shah, M.I.A.; Ateeq, M.; Ullah, M.; Ali, N.; Shaheen, S.; Yasmeen, H.; Shah, S.N.A. Extended visible light driven photocatalytic hydrogen generation by electron induction from g-C3N4 nanosheets to ZnO through the proper heterojunction. Z. Phys. Chem. 2022, 236, 53–66. [Google Scholar] [CrossRef]
  155. Mehtab, A.; Banerjee, S.; Mao, Y.; Ahmad, T. Type-II CuFe2O4/graphitic carbon nitride heterojunctions for high-efficiency photocatalytic and electrocatalytic hydrogen generation. ACS Appl. Mater. Interfaces 2022, 14, 44317–44329. [Google Scholar] [CrossRef]
  156. Mishra, B.P.; Babu, P.; Parida, K. Phosphorous, boron and sulfur doped g-C3N4 nanosheet: Synthesis, characterization, and comparative study towards photocatalytic hydrogen generation. Mater. Today Proc. 2021, 35, 258–262. [Google Scholar] [CrossRef]
  157. Alhaddad, M.; Navarro, R.; Hussein, M.; Mohamed, R. Bi2O3/g-C3N4 nanocomposites as proficient photocatalysts for hydrogen generation from aqueous glycerol solutions beneath visible light. Ceram. Int. 2020, 46, 24873–24881. [Google Scholar] [CrossRef]
  158. Ding, J.; Sun, X.; Wang, Q.; Li, D.-s.; Li, X.; Li, X.; Chen, L.; Zhang, X.; Tian, X.; Ostrikov, K.K. Plasma synthesis of Pt/g-C3N4 photocatalysts with enhanced photocatalytic hydrogen generation. J. Alloys Compd. 2021, 873, 159871. [Google Scholar] [CrossRef]
  159. Jin, Z.; Li, Y.; Hao, X. Ni, Co-based selenide anchored g-C3N4 for boosting photocatalytic hydrogen evolution. Acta Phys.-Chim. Sin. 2021, 37, 1912033. [Google Scholar]
  160. Nagaraja, C.; Kaur, M.; Dhingra, S. Enhanced visible-light-assisted photocatalytic hydrogen generation by MoS2/g-C3N4 nanocomposites. Int. J. Hydrogen Energy 2020, 45, 8497–8506. [Google Scholar] [CrossRef]
  161. Li, J.; Ma, W.; Chen, J.; An, N.; Zhao, Y.; Wang, D.; Mao, Z. Carbon vacancies improved photocatalytic hydrogen generation of g-C3N4 photocatalyst via magnesium vapor etching. Int. J. Hydrogen Energy 2020, 45, 13939–13946. [Google Scholar] [CrossRef]
  162. Ma, X.; Luo, B.; Zeng, Z.; Hu, S.; Jing, D. Significantly Enhanced Photocatalytic Hydrogen Generation over a 2D/2D Z-Scheme La2NiO4/g-C3N4 Hybrid Free of Noble Metal Cocatalyst. ACS Appl. Energy Mater. 2021, 4, 10721–10730. [Google Scholar] [CrossRef]
  163. Che, Y.; Liu, Q.; Lu, B.; Zhai, J.; Wang, K.; Liu, Z. Plasmonic ternary hybrid photocatalyst based on polymeric g-C3N4 towards visible light hydrogen generation. Sci. Rep. 2020, 10, 721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Hou, X.; Cui, L.; Du, H.; Gu, L.; Li, Z.; Yuan, Y. Lowering the schottky barrier of g-C3N4/Carbon graphite heterostructure by N-doping for increased photocatalytic hydrogen generation. Appl. Catal. B Environ. 2020, 278, 119253. [Google Scholar] [CrossRef]
  165. Rawool, S.A.; Samanta, A.; Ajithkumar, T.; Kar, Y.; Polshettiwar, V. Photocatalytic hydrogen generation and CO2 conversion using g-C3N4 decorated dendritic fibrous nanosilica: Role of interfaces between silica and g-C3N4. ACS Appl. Energy Mater. 2020, 3, 8150–8158. [Google Scholar] [CrossRef]
  166. Wan, C.; Zhou, L.; Sun, L.; Xu, L.; Cheng, D.-g.; Chen, F.; Zhan, X.; Yang, Y. Boosting visible-light-driven hydrogen evolution from formic acid over AgPd/2D g-C3N4 nanosheets Mott-Schottky photocatalyst. Chem. Eng. J. 2020, 396, 125229. [Google Scholar] [CrossRef]
  167. Song, T.; Xie, C.; Matras-Postolek, K.; Yang, P. 2D layered g-C3N4/WO3/WS2 S-scheme heterojunctions with enhanced photochemical performance. J. Phys. Chem. C 2021, 125, 19382–19393. [Google Scholar] [CrossRef]
  168. Zhao, W.; She, T.; Zhang, J.; Wang, G.; Zhang, S.; Wei, W.; Yang, G.; Zhang, L.; Xia, D.; Cheng, Z. A novel Z-scheme CeO2/g-C3N4 heterojunction photocatalyst for degradation of Bisphenol A and hydrogen evolution and insight of the photocatalysis mechanism. J. Mater. Sci. Technol. 2021, 85, 18–29. [Google Scholar] [CrossRef]
  169. Liao, Y.; Wang, G.; Wang, J.; Wang, K.; Yan, S.; Su, Y. Nitrogen vacancy induced in situ g-C3N4 pn homojunction for boosting visible light-driven hydrogen evolution. J. Colloid Interface Sci. 2021, 587, 110–120. [Google Scholar] [CrossRef]
  170. Zhang, X.; Zhang, X.R.; Yang, P.; Chen, H.-S. Black magnetic Cu-g-C3N4 nanosheets towards efficient photocatalytic H2 generation and CO2/benzene conversion. Chem. Eng. J. 2022, 450, 138030. [Google Scholar] [CrossRef]
  171. Zhao, B.; Xu, J.; Liu, Y.; Fan, J.; Yu, H. Amino group-rich porous g-C3N4 nanosheet photocatalyst: Facile oxalic acid-induced synthesis and improved H2-evolution activity. Ceram. Int. 2021, 47, 18295–18303. [Google Scholar] [CrossRef]
  172. Yuan, H.; Fang, F.; Dong, J.; Xia, W.; Zeng, X.; Shangguan, W. Enhanced photocatalytic hydrogen production based on laminated MoS2/g-C3N4 photocatalysts. Colloids Surf. A Physicochem. Eng. Asp. 2022, 641, 128575. [Google Scholar] [CrossRef]
  173. Kamat, P.V. Meeting the clean energy demand: Nanostructure architectures for solar energy conversion. J. Phys. Chem. C 2007, 111, 2834–2860. [Google Scholar] [CrossRef]
  174. Wang, Q.; Fang, Z.; Zhang, W.; Zhang, D. High-efficiency g-C3N4 based photocatalysts for CO2 reduction: Modification methods. Adv. Fiber Mater. 2022, 4, 342–360. [Google Scholar] [CrossRef]
  175. Fu, J.; Yu, J.; Jiang, C.; Cheng, B. g-C3N4-based heterostructured photocatalysts. Adv. Energy Mater. 2018, 8, 1701503. [Google Scholar] [CrossRef]
  176. Kumar, S.; Karthikeyan, S.; Lee, A.F. g-C3N4-based nanomaterials for visible light-driven photocatalysis. Catalysts 2018, 8, 74. [Google Scholar] [CrossRef] [Green Version]
  177. Li, Y.; Kong, T.; Shen, S. Artificial photosynthesis with polymeric carbon nitride: When meeting metal nanoparticles, single atoms, and molecular complexes. Small 2019, 15, 1900772. [Google Scholar] [CrossRef]
  178. Zhang, L.; Long, R.; Zhang, Y.; Duan, D.; Xiong, Y.; Zhang, Y.; Bi, Y. Direct observation of dynamic bond evolution in single-atom Pt/C3N4 catalysts. Angew. Chem. 2020, 132, 6283–6288. [Google Scholar] [CrossRef]
  179. Liu, W.; Cao, L.; Cheng, W.; Cao, Y.; Liu, X.; Zhang, W.; Mou, X.; Jin, L.; Zheng, X.; Che, W. Single-site active cobalt-based photocatalyst with a Long carrier lifetime for spontaneous overall water splitting. Angew. Chem. 2017, 129, 9440–9445. [Google Scholar] [CrossRef]
  180. Pan, Z.; Zheng, Y.; Guo, F.; Niu, P.; Wang, X. Decorating CoP and Pt nanoparticles on graphitic carbon nitride nanosheets to promote overall water splitting by conjugated polymers. ChemSusChem 2017, 10, 87–90. [Google Scholar] [CrossRef]
  181. Reddy, K.R.; Reddy, C.V.; Nadagouda, M.N.; Shetti, N.P.; Jaesool, S.; Aminabhavi, T.M. Polymeric graphitic carbon nitride (g-C3N4)-based semiconducting nanostructured materials: Synthesis methods, properties and photocatalytic applications. J. Environ. Manag. 2019, 238, 25–40. [Google Scholar] [CrossRef] [PubMed]
  182. Wen, J.; Xie, J.; Chen, X.; Li, X. A review on g-C3N4-based photocatalysts. Appl. Surf. Sci. 2017, 391, 72–123. [Google Scholar] [CrossRef]
  183. Wang, K.; Li, Q.; Liu, B.; Cheng, B.; Ho, W.; Yu, J. Sulfur-doped g-C3N4 with enhanced photocatalytic CO2-reduction performance. Appl. Catal. B Environ. 2015, 176, 44–52. [Google Scholar] [CrossRef]
  184. Sun, Z.; Wang, H.; Wu, Z.; Wang, L. g-C3N4 based composite photocatalysts for photocatalytic CO2 reduction. Catal. Today 2018, 300, 160–172. [Google Scholar] [CrossRef]
  185. Ghosh, U.; Majumdar, A.; Pal, A. Photocatalytic CO2 reduction over g-C3N4 based heterostructures: Recent progress and prospects. J. Environ. Chem. Eng. 2021, 9, 104631. [Google Scholar] [CrossRef]
  186. Li, X.; Xiong, J.; Gao, X.; Huang, J.; Feng, Z.; Chen, Z.; Zhu, Y. Recent advances in 3D g-C3N4 composite photocatalysts for photocatalytic water splitting, degradation of pollutants and CO2 reduction. J. Alloys Compd. 2019, 802, 196–209. [Google Scholar] [CrossRef]
  187. Zhang, X. Layered g-C3N4/TiO2 nanocomposites for efficient photocatalytic water splitting and CO2 reduction: A review. Mater. Today Energy 2022, 23, 100904. [Google Scholar] [CrossRef]
  188. Chen, P.; Dong, X.a.; Huang, M.; Li, K.; Xiao, L.; Sheng, J.; Chen, S.; Zhou, Y.; Dong, F. Rapid self-decomposition of g-C3N4 during gas–solid photocatalytic CO2 reduction and its effects on performance assessment. ACS Catal. 2022, 12, 4560–4570. [Google Scholar] [CrossRef]
  189. Li, H.; Wang, D.; Miao, C.; Wang, Y.; Wang, Y.; Liu, C.; Che, G. g-C3N4/BiOI S-scheme heterojunction: A 2D/2D model platform for visible-light-driven photocatalytic CO2 reduction and pollutant degradation. J. Environ. Chem. Eng. 2022, 10, 108201. [Google Scholar] [CrossRef]
  190. Fu, J.; Zhu, B.; Jiang, C.; Cheng, B.; You, W.; Yu, J. Hierarchical porous O-doped g-C3N4 with enhanced photocatalytic CO2 reduction activity. Small 2017, 13, 1603938. [Google Scholar] [CrossRef]
  191. Huo, Y.; Zhang, J.; Dai, K.; Liang, C. Amine-modified S-scheme porous g-C3N4/CdSe–diethylenetriamine composite with enhanced photocatalytic CO2 reduction activity. ACS Appl. Energy Mater. 2021, 4, 956–968. [Google Scholar] [CrossRef]
  192. Tang, R.; Wang, H.; Zhang, S.; Zhang, L.; Dong, F. A ball milling method for highly dispersed Ni atoms on g-C3N4 to boost CO2 photoreduction. J. Colloid Interface Sci. 2023, 630, 290–300. [Google Scholar] [CrossRef] [PubMed]
  193. Lu, Z.; Wang, Z. S-scheme CuWO4@ g-C3N4 core-shell microsphere for CO2 photoreduction. Mater. Sci. Semicond. Process. 2023, 153, 107177. [Google Scholar] [CrossRef]
  194. Li, J.; He, C.; Xu, N.; Wu, K.; Huang, Z.; Zhao, X.; Nan, J.; Xiao, X. Interfacial bonding of hydroxyl-modified g-C3N4 and Bi2O2CO3 toward boosted CO2 photoreduction: Insights into the key role of OH groups. Chem. Eng. J. 2023, 452, 139191. [Google Scholar] [CrossRef]
  195. Gao, Y.; Sun, L.; Bian, J.; Zhang, Z.; Li, Z.; Jing, L. Accelerated charge transfer of g-C3N4/BiVO4 Z-scheme 2D heterojunctions by controllably introducing phosphate bridges and Ag nanocluster co-catalysts for selective CO2 photoreduction to CO. Appl. Surf. Sci. 2023, 610, 155360. [Google Scholar] [CrossRef]
  196. Wang, K.; Feng, X.; Shangguan, Y.; Wu, X.; Chen, H. Selective CO2 photoreduction to CH4 mediated by dimension-matched 2D/2D Bi3NbO7/g-C3N4 S-scheme heterojunction. Chin. J. Catal. 2022, 43, 246–254. [Google Scholar] [CrossRef]
  197. Wang, J.; Yu, Y.; Cui, J.; Li, X.; Zhang, Y.; Wang, C.; Yu, X.; Ye, J. Defective g-C3N4/covalent organic framework van der Waals heterojunction toward highly efficient S-scheme CO2 photoreduction. Appl. Catal. B Environ. 2022, 301, 120814. [Google Scholar] [CrossRef]
  198. Tao, K.-Y.; Yuan, K.; Yang, W.; Zhong, D.-C.; Lu, T.-B. A template co-pyrolysis strategy towards the increase of amino/imino content within g-C3N4 for efficient CO2 photoreduction. Chem. Eng. J. 2022, 140630. [Google Scholar] [CrossRef]
  199. Tang, Z.; Wang, C.; He, W.; Wei, Y.; Zhao, Z.; Liu, J. The Z-scheme g-C3N4/3DOM-WO3 photocatalysts with enhanced activity for CO2 photoreduction into CO. Chin. Chem. Lett. 2022, 33, 939–942. [Google Scholar] [CrossRef]
  200. Sahoo, R.C.; Lu, H.; Garg, D.; Yin, Z.; Matte, H.R. Bandgap engineered g-C3N4 and its graphene composites for stable photoreduction of CO2 to methanol. Carbon 2022, 192, 101–108. [Google Scholar] [CrossRef]
  201. Madi, M.; Tahir, M. Highly stable LaCoO3 perovskite supported g-C3N4 nanotextures with proficient charges migration for visible light CO2 photoreduction to CO and CH4. Mater. Sci. Semicond. Process. 2022, 142, 106517. [Google Scholar] [CrossRef]
  202. Liu, T.; Hao, L.; Bai, L.; Liu, J.; Zhang, Y.; Tian, N.; Huang, H. Z-scheme junction Bi2O2 (NO3)(OH)/g-C3N4 for promoting CO2 photoreduction. Chem. Eng. J. 2022, 429, 132268. [Google Scholar] [CrossRef]
  203. Li, Y.; Yin, Q.; Zeng, Y.; Liu, Z. Hollow spherical biomass derived-carbon dotted with SnS2/g-C3N4 Z-scheme heterojunction for efficient CO2 photoreduction into CO. Chem. Eng. J. 2022, 438, 135652. [Google Scholar] [CrossRef]
  204. Li, X.; Li, S.; Xu, J.; Wang, L.; Liang, K.; Zhang, H.; Yu, X.; Liu, Z. Synergy of nitrogen vacancies and nanodiamond decoration in g-C3N4 for boosting CO2 photoreduction. Appl. Surf. Sci. 2022, 600, 154199. [Google Scholar] [CrossRef]
  205. Li, L.; Ma, D.; Xu, Q.; Huang, S. Constructing hierarchical ZnIn2S4/g-C3N4 S-scheme heterojunction for boosted CO2 photoreduction performance. Chem. Eng. J. 2022, 437, 135153. [Google Scholar] [CrossRef]
  206. He, J.; Wang, X.; Jin, S.; Liu, Z.-Q.; Zhu, M. 2D metal-free heterostructure of covalent triazine framework/g-C3N4 for enhanced photocatalytic CO2 reduction with high selectivity. Chin. J. Catal. 2022, 43, 1306–1315. [Google Scholar] [CrossRef]
  207. Zhu, X.; Deng, H.; Cheng, G. Facile construction of g-C3N4-W18O49 heterojunction with improved charge transfer for solar-driven CO2 photoreduction. Inorg. Chem. Commun. 2021, 132, 108814. [Google Scholar] [CrossRef]
  208. Yin, S.; Sun, L.; Zhou, Y.; Li, X.; Li, J.; Song, X.; Huo, P.; Wang, H.; Yan, Y. Enhanced electron–hole separation in SnS2/Au/g-C3N4 embedded structure for efficient CO2 photoreduction. Chem. Eng. J. 2021, 406, 126776. [Google Scholar] [CrossRef]
  209. Xu, Y.; Jin, X.; Ge, T.; Xie, H.; Sun, R.; Su, F.; Li, X.; Ye, L. Realizing efficient CO2 photoreduction in Bi3O4Cl: Constructing van der Waals heterostructure with g-C3N4. Chem. Eng. J. 2021, 409, 128178. [Google Scholar] [CrossRef]
  210. She, S.; Zhang, X.; Wu, X.; Li, J.; Zhang, G. The fabrication of two-dimensional g-C3N4/NaBiO3· 2H2O heterojunction for improved photocatalytic CO2 reduction: DFT study and mechanism unveiling. J. Colloid Interface Sci. 2021, 604, 122–130. [Google Scholar] [CrossRef]
  211. Qin, Y.; Dong, G.; Zhang, L.; Li, G.; An, T. Highly efficient and selective photoreduction of CO2 to CO with nanosheet g-C3N4 as compared with its bulk counterpart. Environ. Res. 2021, 195, 110880. [Google Scholar] [CrossRef] [PubMed]
  212. Mkhalid, I.A.; Mohamed, R.M.; Ismail, A.A.; Alhaddad, M. Z-scheme g-C3N4 nanosheet photocatalyst decorated with mesoporous CdS for the photoreduction of carbon dioxide. Ceram. Int. 2021, 47, 17210–17219. [Google Scholar] [CrossRef]
  213. Li, X.; Jiang, H.; Ma, C.; Zhu, Z.; Song, X.; Wang, H.; Huo, P.; Li, X. Local surface plasma resonance effect enhanced Z-scheme ZnO/Au/g-C3N4 film photocatalyst for reduction of CO2 to CO. Appl. Catal. B Environ. 2021, 283, 119638. [Google Scholar] [CrossRef]
  214. Li, X.; Guan, J.; Jiang, H.; Song, X.; Huo, P.; Wang, H. rGO modified R-CeO2/g-C3N4 multi-interface contact S-scheme photocatalyst for efficient CO2 photoreduction. Appl. Surf. Sci. 2021, 563, 150042. [Google Scholar] [CrossRef]
  215. Chen, C.; Jin, J.; Chen, S.; Wang, T.; Xiao, J.; Peng, T. In-situ growth of ultrafine ZnO on g-C3N4 layer for highly active and selective CO2 photoreduction to CH4 under visible light. Mater. Res. Bull. 2021, 137, 111177. [Google Scholar] [CrossRef]
  216. Wang, S.; Zhan, J.; Chen, K.; Ali, A.; Zeng, L.; Zhao, H.; Hu, W.; Zhu, L.; Xu, X. Potassium-doped g-C3N4 achieving efficient visible-light-driven CO2 reduction. ACS Sustain. Chem. Eng. 2020, 8, 8214–8222. [Google Scholar] [CrossRef]
  217. Vu, N.-N.; Kaliaguine, S.; Do, T.-O. Synthesis of the g-C3N4/CdS nanocomposite with a chemically bonded interface for enhanced sunlight-driven CO2 photoreduction. ACS Appl. Energy Mater. 2020, 3, 6422–6433. [Google Scholar] [CrossRef]
  218. Ma, W.; Wang, N.; Guo, Y.; Yang, L.; Lv, M.; Tang, X.; Li, S. Enhanced photoreduction CO2 activity on g-C3N4: By synergistic effect of nitrogen defective-enriched and porous structure, and mechanism insights. Chem. Eng. J. 2020, 388, 124288. [Google Scholar] [CrossRef]
  219. Guo, L.; You, Y.; Huang, H.; Tian, N.; Ma, T.; Zhang, Y. Z-scheme g-C3N4/Bi2O2 [BO2 (OH)] heterojunction for enhanced photocatalytic CO2 reduction. J. Colloid Interface Sci. 2020, 568, 139–147. [Google Scholar] [CrossRef]
  220. Guo, H.; Ding, J.; Wan, S.; Wang, Y.; Zhong, Q. Highly efficient CH3OH production over Zn0. 2Cd0. 8S decorated g-C3N4 heterostructures for the photoreduction of CO2. Appl. Surf. Sci. 2020, 528, 146943. [Google Scholar] [CrossRef]
  221. Khan, I.; Yuan, A.; Khan, S.; Khan, A.; Khan, S.; Shah, S.A.; Luo, M.; Yaseen, W.; Shen, X.; Yaseen, M. Graphitic carbon nitride composites with gold and ZIF-67 nanoparticles as visible-light-promoted catalysts for CO2 conversion and bisphenol A degradation. ACS Appl. Nano Mater. 2022, 5, 13404–13416. [Google Scholar] [CrossRef]
  222. Dao, X.-Y.; Xie, X.-F.; Guo, J.-H.; Zhang, X.-Y.; Kang, Y.-S.; Sun, W.-Y. Boosting photocatalytic CO2 reduction efficiency by heterostructures of NH2-MIL-101 (Fe)/g-C3N4. ACS Appl. Energy Mater. 2020, 3, 3946–3954. [Google Scholar] [CrossRef]
  223. Nemiwal, M.; Zhang, T.C.; Kumar, D. Recent progress in g-C3N4, TiO2 and ZnO based photocatalysts for dye degradation: Strategies to improve photocatalytic activity. Sci. Total Environ. 2021, 767, 144896. [Google Scholar] [CrossRef] [PubMed]
  224. Duan, Y.; Deng, L.; Shi, Z.; Liu, X.; Zeng, H.; Zhang, H.; Crittenden, J. Efficient sulfadiazine degradation via in-situ epitaxial grow of graphitic carbon nitride (g-C3N4) on carbon dots heterostructures under visible light irradiation: Synthesis, mechanisms and toxicity evaluation. J. Colloid Interface Sci. 2020, 561, 696–707. [Google Scholar] [CrossRef] [PubMed]
  225. Zhao, Y.; Shi, H.; Yang, D.; Fan, J.; Hu, X.; Liu, E. Fabrication of a Sb2MoO6/g-C3N4 photocatalyst for enhanced RhB degradation and H2 generation. J. Phys. Chem. C 2020, 124, 13771–13778. [Google Scholar] [CrossRef]
  226. Hayat, A.; Sohail, M.; Ali Shah Syed, J.; Al-Sehemi, A.G.; Mohammed, M.H.; Al-Ghamdi, A.A.; Taha, T.; Salem AlSalem, H.; Alenad, A.M.; Amin, M.A. Recent advancement of the current aspects of g-C3N4 for its photocatalytic applications in sustainable energy system. Chem. Rec. 2022, 22, e202100310. [Google Scholar] [CrossRef] [PubMed]
  227. Li, J.; Qi, Y.; Mei, Y.; Ma, S.; Li, Q.; Xin, B.; Yao, T.; Wu, J. Construction of phosphorus-doped carbon nitride/phosphorus and sulfur co-doped carbon nitride isotype heterojunction and their enhanced photoactivity. J. Colloid Interface Sci. 2020, 566, 495–504. [Google Scholar] [CrossRef]
  228. Zhang, B.; Li, X.; Zhao, Y.; Song, H.; Wang, H. Facile synthesis of oxygen doped mesoporous graphitic carbon nitride with high photocatalytic degradation efficiency under simulated solar irradiation. Colloids Surf. A Physicochem. Eng. Asp. 2019, 580, 123736. [Google Scholar] [CrossRef]
  229. Jiang, L.; Yuan, X.; Zeng, G.; Liang, J.; Wu, Z.; Wang, H.; Zhang, J.; Xiong, T.; Li, H. A facile band alignment of polymeric carbon nitride isotype heterojunctions for enhanced photocatalytic tetracycline degradation. Environ. Sci. Nano 2018, 5, 2604–2617. [Google Scholar] [CrossRef]
  230. Chen, M.; Jia, Y.; Li, H.; Wu, Z.; Huang, T.; Zhang, H. Enhanced pyrocatalysis of the pyroelectric BiFeO3/g-C3N4 heterostructure for dye decomposition driven by cold-hot temperature alternation. J. Adv. Ceram. 2021, 10, 338–346. [Google Scholar] [CrossRef]
  231. Zhang, H.; Guo, L.-H.; Zhao, L.; Wan, B.; Yang, Y. Switching oxygen reduction pathway by exfoliating graphitic carbon nitride for enhanced photocatalytic phenol degradation. J. Phys. Chem. Lett. 2015, 6, 958–963. [Google Scholar] [CrossRef] [PubMed]
  232. Li, X.; Yu, J.; Jaroniec, M. Hierarchical photocatalysts. Chem. Soc. Rev. 2016, 45, 2603–2636. [Google Scholar] [CrossRef] [PubMed]
  233. Du, X.; Bai, X.; Xu, L.; Yang, L.; Jin, P. Visible-light activation of persulfate by TiO2/g-C3N4 photocatalyst toward efficient degradation of micropollutants. Chem. Eng. J. 2020, 384, 123245. [Google Scholar] [CrossRef]
  234. Zhou, W.-J.; Ma, L.-X.; Li, L.-Y.; Zha, M.; Li, B.-L.; Wu, B.; Hu, C.-J. Synthesis of a 3D Cu (II) MOF and its heterostructual g-C3N4 composite showing improved visible-light-driven photodegradation of organic dyes. J. Solid State Chem. 2022, 315, 123520. [Google Scholar] [CrossRef]
  235. Shekardasht, M.B.; Givianrad, M.H.; Gharbani, P.; Mirjafary, Z.; Mehrizad, A. Preparation of a novel Z-scheme g-C3N4/RGO/Bi2Fe4O9 nanophotocatalyst for degradation of Congo Red dye under visible light. Diam. Relat. Mater. 2020, 109, 108008. [Google Scholar] [CrossRef]
  236. Liu, Y.; Shen, Z.; Song, J.; Qi, H.; Yin, C.; Zou, X.; Xia, Q.; Cui, L.; Li, X.; Wang, Y. Honeycomb-like g-C3N4/CeO2-x nanosheets obtained via one step hydrothermal-roasting for efficient and stable Cr (VI) photo-reduction. Chin. Chem. Lett. 2020, 31, 2747–2751. [Google Scholar] [CrossRef]
  237. Prabavathi, S.L.; Saravanakumar, K.; Park, C.M.; Muthuraj, V. Photocatalytic degradation of levofloxacin by a novel Sm6WO12/g-C3N4 heterojunction: Performance, mechanism and degradation pathways. Sep. Purif. Technol. 2021, 257, 117985. [Google Scholar] [CrossRef]
  238. Deng, Y.; Liu, J.; Huang, Y.; Ma, M.; Liu, K.; Dou, X.; Wang, Z.; Qu, S.; Wang, Z. Engineering the photocatalytic behaviors of g/C3N4-based metal-free materials for degradation of a representative antibiotic. Adv. Funct. Mater. 2020, 30, 2002353. [Google Scholar] [CrossRef]
  239. Paul, D.R.; Gautam, S.; Panchal, P.; Nehra, S.P.; Choudhary, P.; Sharma, A. ZnO-modified g-C3N4: A potential photocatalyst for environmental application. ACS Omega 2020, 5, 3828–3838. [Google Scholar] [CrossRef] [Green Version]
  240. Wu, L.; Chen, Y.; Li, Y.; Meng, Q.; Duan, T. Functionally integrated g-C3N4@ wood-derived carbon with an orderly interconnected porous structure. Appl. Surf. Sci. 2021, 540, 148440. [Google Scholar] [CrossRef]
  241. Yassin, J.M.; Taddesse, A.M.; Sánchez-Sánchez, M. Sustainable synthesis of semicrystalline Zr-BDC MOF and heterostructural Ag3PO4/Zr-BDC/g-C3N4 composite for photocatalytic dye degradation. Catal. Today 2022, 390, 162–175. [Google Scholar] [CrossRef]
  242. Hasija, V.; Raizada, P.; Sudhaik, A.; Singh, P.; Thakur, V.K.; Khan, A.A.P. Fabrication of Ag/AgI/WO3 heterojunction anchored P and S co-doped graphitic carbon nitride as a dual Z scheme photocatalyst for efficient dye degradation. Solid State Sci. 2020, 100, 106095. [Google Scholar] [CrossRef]
  243. Feng, Z.; Zeng, L.; Zhang, Q.; Ge, S.; Zhao, X.; Lin, H.; He, Y. In situ preparation of g-C3N4/Bi4O5I2 complex and its elevated photoactivity in Methyl Orange degradation under visible light. J. Environ. Sci. 2020, 87, 149–162. [Google Scholar] [CrossRef] [PubMed]
  244. Zheng, C.-W.; Niu, H.-Y.; Liang, C.; Niu, C.-G.; Zhao, X.-F.; Zhang, L.; Li, J.-S.; Guo, H.; Liu, H.-Y.; Liang, S. A study on advanced oxidation mechanism of MnCo2O4/g-C3N4 degradation of nitrobenzene: Sacrificial oxidation and radical oxidation. Chem. Eng. J. 2021, 403, 126400. [Google Scholar] [CrossRef]
  245. Yu, X.; Zhang, X.; Zhao, J.; Xu, L.; Yan, J. Flower-like shaped Bi12TiO20/g-C3N4 heterojunction for effective elimination of organic pollutants: Preparation, characterization, and mechanism study. Appl. Organomet. Chem. 2020, 34, e5702. [Google Scholar] [CrossRef]
  246. Feng, S.; Chen, T.; Liu, Z.; Shi, J.; Yue, X.; Li, Y. Z-scheme CdS/CQDs/g-C3N4 composites with visible-near-infrared light response for efficient photocatalytic organic pollutant degradation. Sci. Total Environ. 2020, 704, 135404. [Google Scholar] [CrossRef]
  247. Liang, Q.; Liu, X.; Shao, B.; Tang, L.; Liu, Z.; Zhang, W.; Gong, S.; Liu, Y.; He, Q.; Wu, T. Construction of fish-scale tubular carbon nitride-based heterojunction with boosting charge separation in photocatalytic tetracycline degradation and H2O2 production. Chem. Eng. J. 2021, 426, 130831. [Google Scholar] [CrossRef]
  248. Huang, Y.; Li, B.; Wu, F.; Yang, B. Fabrication of novel flower-like Co3O4/g-C3N4 heterojunction for tetracycline degradation under visible light irradiation. Mater. Lett. 2022, 311, 131538. [Google Scholar] [CrossRef]
  249. Chen, W.; Huang, J.; He, Z.-C.; Ji, X.; Zhang, Y.-F.; Sun, H.-L.; Wang, K.; Su, Z.-W. Accelerated photocatalytic degradation of tetracycline hydrochloride over CuAl2O4/g-C3N4 pn heterojunctions under visible light irradiation. Sep. Purif. Technol. 2021, 277, 119461. [Google Scholar] [CrossRef]
  250. Shi, Y.; Li, J.; Wan, D.; Huang, J.; Liu, Y. Peroxymonosulfate-enhanced photocatalysis by carbonyl-modified g-C3N4 for effective degradation of the tetracycline hydrochloride. Sci. Total Environ. 2020, 749, 142313. [Google Scholar] [CrossRef]
  251. Luo, J.; Dai, Y.; Xu, X.; Liu, Y.; Yang, S.; He, H.; Sun, C.; Xian, Q. Green and efficient synthesis of Co-MOF-based/g-C3N4 composite catalysts to activate peroxymonosulfate for degradation of the antidepressant venlafaxine. J. Colloid Interface Sci. 2022, 610, 280–294. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures of (a) triazine and (b) tri-s-triazine. Reprinted with permission from Ref. [25]. Copyright 2021 American Chemical Society.
Figure 1. Structures of (a) triazine and (b) tri-s-triazine. Reprinted with permission from Ref. [25]. Copyright 2021 American Chemical Society.
Molecules 28 00432 g001
Figure 2. Illustration of HCNS and metal/HCNS composite syntheses. Reprinted with permission from Ref. [34]. Copyright 2012 Nature Publishing Group.
Figure 2. Illustration of HCNS and metal/HCNS composite syntheses. Reprinted with permission from Ref. [34]. Copyright 2012 Nature Publishing Group.
Molecules 28 00432 g002
Figure 3. Schematic illustration of different techniques for the synthesis of (a) 0D g-C3N4, (b) 1D g-C3N4, (c) 2D g-C3N4, and (d) 3D g-C3N4. Reprinted with permission from Ref. [25]. Copyright 2021 American Chemical Society.
Figure 3. Schematic illustration of different techniques for the synthesis of (a) 0D g-C3N4, (b) 1D g-C3N4, (c) 2D g-C3N4, and (d) 3D g-C3N4. Reprinted with permission from Ref. [25]. Copyright 2021 American Chemical Society.
Molecules 28 00432 g003
Figure 4. XPS spectra of samples pure 650 CN and Fe3+ R-650 CN. (a) C 1s. (b) N 1s. (c) O 1s. (d) Fe 2p. Reprinted with permission from Ref. [49]. Copyright 2019 John Wiley and Sons.
Figure 4. XPS spectra of samples pure 650 CN and Fe3+ R-650 CN. (a) C 1s. (b) N 1s. (c) O 1s. (d) Fe 2p. Reprinted with permission from Ref. [49]. Copyright 2019 John Wiley and Sons.
Molecules 28 00432 g004
Figure 5. UV-visible absorption spectra (A) and fluorescence emission spectra (B) of: (a) bulk g-C3N4 and (b) g-C3N4 nanosheets. The wavelength of excitation light for fluorescence emission spectra is 350 nm. Reprinted with permission from Ref. [52]. Copyright 2012 John Wiley and Sons.
Figure 5. UV-visible absorption spectra (A) and fluorescence emission spectra (B) of: (a) bulk g-C3N4 and (b) g-C3N4 nanosheets. The wavelength of excitation light for fluorescence emission spectra is 350 nm. Reprinted with permission from Ref. [52]. Copyright 2012 John Wiley and Sons.
Molecules 28 00432 g005
Figure 6. SEM images of hollow mesoporous carbon nitride spheres with different content of ionic liquid. Reprinted with permission from Ref. [35]. Copyright 2018 Elsevier.
Figure 6. SEM images of hollow mesoporous carbon nitride spheres with different content of ionic liquid. Reprinted with permission from Ref. [35]. Copyright 2018 Elsevier.
Molecules 28 00432 g006
Figure 7. Illustration of the fabrication of porous O-doped g-C3N4 from hydrogen bond-induced supramolecular precursor assembled under hydrothermal treatment. Reprinted with permission from Ref. [64]. Copyright 2018 Elsevier.
Figure 7. Illustration of the fabrication of porous O-doped g-C3N4 from hydrogen bond-induced supramolecular precursor assembled under hydrothermal treatment. Reprinted with permission from Ref. [64]. Copyright 2018 Elsevier.
Molecules 28 00432 g007
Figure 8. Schematic illustration of the synthesis process of Fe-doped g-C3N4 nanosheets. Reprinted with permission from Ref. [68]. Copyright 2017 John Wiley and Sons.
Figure 8. Schematic illustration of the synthesis process of Fe-doped g-C3N4 nanosheets. Reprinted with permission from Ref. [68]. Copyright 2017 John Wiley and Sons.
Molecules 28 00432 g008
Figure 9. Schematic illustration of enhancement mechanism over carbon-induced g-C3N4 nanocomposites in photocatalytic action. (a) Different types of carbon materials. (b) Chemical structure of g-C3N4; black and red dots represent C and N, respectively. (c) The enhanced photocatalytic mechanisms and (d) energetic photocatalytic application of carbon-induced metal-free g-C3N4 nanocomposites. Reprinted with permission from Ref. [121]. Copyright 2020 John Wiley and Sons.
Figure 9. Schematic illustration of enhancement mechanism over carbon-induced g-C3N4 nanocomposites in photocatalytic action. (a) Different types of carbon materials. (b) Chemical structure of g-C3N4; black and red dots represent C and N, respectively. (c) The enhanced photocatalytic mechanisms and (d) energetic photocatalytic application of carbon-induced metal-free g-C3N4 nanocomposites. Reprinted with permission from Ref. [121]. Copyright 2020 John Wiley and Sons.
Molecules 28 00432 g009
Figure 10. Schematic illustration of the preparation process for g-C3N4/CD-based nanocomposites through different methods. (a) Mechanical mixing method. (b) Ultrasonication method. (c) Electrostatic attraction method. (d) Hydrothermal/solvothermal method. (e,f) Calcine method. Reprinted with permission from Ref. [130]. Copyright 2021 John Wiley and Sons.
Figure 10. Schematic illustration of the preparation process for g-C3N4/CD-based nanocomposites through different methods. (a) Mechanical mixing method. (b) Ultrasonication method. (c) Electrostatic attraction method. (d) Hydrothermal/solvothermal method. (e,f) Calcine method. Reprinted with permission from Ref. [130]. Copyright 2021 John Wiley and Sons.
Molecules 28 00432 g010
Figure 11. Scheme for a one-pot thermal preparation of amino-functionalized ultrathin nanoporous B-doped g-C3N4. Reprinted with permission from Ref. [138]. Copyright 2021 Elsevier.
Figure 11. Scheme for a one-pot thermal preparation of amino-functionalized ultrathin nanoporous B-doped g-C3N4. Reprinted with permission from Ref. [138]. Copyright 2021 Elsevier.
Molecules 28 00432 g011
Figure 12. Schematic diagram for the performance comparison on hydrogen evolution between metal- and non-metal-modified g-C3N4 composites. Reprinted with permission from Ref. [140]. Copyright 2023 Elsevier.
Figure 12. Schematic diagram for the performance comparison on hydrogen evolution between metal- and non-metal-modified g-C3N4 composites. Reprinted with permission from Ref. [140]. Copyright 2023 Elsevier.
Molecules 28 00432 g012
Figure 13. Schematic illustration of energy levels of PCN for photocatalytic CO2 reduction. Reprinted with permission from Ref. [177]. Copyright 2019 John Wiley and Sons.
Figure 13. Schematic illustration of energy levels of PCN for photocatalytic CO2 reduction. Reprinted with permission from Ref. [177]. Copyright 2019 John Wiley and Sons.
Molecules 28 00432 g013
Figure 14. Photocatalytic mechanism of g-C3N4/rGO (reduced graphene oxide) for the degradation of Rhodamine B. Reprinted with permission from Ref. [122]. Copyright 2015 Elsevier.
Figure 14. Photocatalytic mechanism of g-C3N4/rGO (reduced graphene oxide) for the degradation of Rhodamine B. Reprinted with permission from Ref. [122]. Copyright 2015 Elsevier.
Molecules 28 00432 g014
Table 1. Photocatalytic H2 generation over g-C3N4-based materials.
Table 1. Photocatalytic H2 generation over g-C3N4-based materials.
EntryPhotocatalystExperimental
Details
H2 Evolution RateReference
Material
/μmol·g−1·h−1
Enhancement Relative to
Conventional g-C3N4
Apparent Quantum
Efficiency/%
Ref.
1Cyano-group-modified crystalline g-C3N4 (CCN0.1)50 mg CCN, 80 mL lactic acid (10 vol%), 1 wt% Pt/CCN 758.8 μmol·h−1Bulk g-C3N4
379.4 μmol·h−1
21.17%[142]
2Ba5Nb4O15/g-C3N4 (1:20)420 nm LEDs (3 W single), Pt co-catalyst, 0.05 g, 100 mL oxalic acid 2.67 μmol·h−1g-C3N4
1.14 μmol·h−1
2.356.1[143]
3P-doped g-C3N4 with aromatic ring (AS/P-CN)300 W Xe, 10 vol% TEOA, 2 wt% Pt, 40 mg550 μmol·h−1·g−1Pristine CN
120 μmol·h−1·g−1
4.580.33[144]
4Ni3S2-NiS2/CN-3300 W Xe, 20 mg Ni3S2-NiS2/CN-X, 90 mL water, 10 mL TEOA1206.6 μmol·h−1·g−1Pure g-C3N4
4.01 μmol·h−1·g−1
300.7 [145]
52% MoS2-g-C3N4/Ni2P300 W Xe (>420 nm), 5 °C, 50 mg, 90 mL water, 10 mL of TEOA298.1 μmol·h−1·g−1Pure g-C3N4
4.32 μmol·h−1·g−1
692.51% (λ = 420 nm)[146]
6HBTiO2/g-C3N4 QDs300 W Xe,
0.025 g, 50 mL 0.25 M Na2S, and 0.35M Na2SO3
10.57 mmol h−1·g−1g-C3N4
0.32 mmol. h−1·g−1
3318.6%
420 nm
[147]
70.8 wt.% g-C3N4/BiVO4200 mL lake water, 500 W halogen, 0.5 M Na2SO421.4 mmol h−1 4.27% at 420 nm[148]
8g-C3N4/CoP-4%350 W Xe, 10 mg, 70 mL water, 10 mL TEOA936 μmol g−1 h−1g-C3N4—4 wt% Pt 665 μmol g−1 h−11.41 [149]
9NiCoP-3/C3N4300 W Xe (300–780 nm), 100 mg, 10 mL methanol, 90 mL water159 μmol g−1 h−1CoP-3/C3N4: 63.6 μmol g−1 h−1
Ni2P-3/C3N4: 4.54 μmol g−1 h−1
CoP-3/C3N4: 2.5
Ni2P-3/C3N4: 35
4.2%[150]
1015% FeSe2/CNNS 2D/2D composite0.15/0.35 mol/L Na2S/Na2SO3,
30 mg, 300 W Xe
1655.6 μmol g−1 h−1C3N4: 624.8 μmol g−1 h−1
FeSe2: 957 μmol g−1 h−1
Pristine g-C3N4: 2.65
FeSe2: 1.73
[151]
115%-NiCo2O4/g-C3N4300 W Xe, 50 mg, 100 mL
solution (10 vol% of TEOA, 3%
of H2PtCl6)
1041.9 μmol g−1 h−1g-C3N4
521.4 μmol g−1 h−1
2 [152]
12CoS/g-C3N4/NiS
ternary photocatalyst
300 W Xe, 100 mg, 85 mL water, 15 mL TEOA1.93 mmol h−1·g−1Bare g-C3N4
0.15 mmol h−1·g−1
12.816.4% at 420 nm[153]
135ZnO/
g-C3N4
0.2 g, 80 mL deionized water, 20 mL methanol70 µmol h−1g-C3N4
8 µmol h−1
8.75 [154]
1420 wt% CuFe2O4/g-C3N4200 W Hg−Xe, 20 mg, Na2S/Na2SO3, TEOA700.34 μmol g−1 h−1g-C3N4 nanosheets
280.1 μmol g−1 h−1
2.525.09%[155]
15Boron-doped g-C3N4150 W Xe, 20 mg, 10 vol% TEOA, Pt (1 wt%)18.2 µmol h−1g-C3N4
6.1 µmol h−1
3 [156]
163.0% β-Bi2O3/g-C3N4500 W Xe, 50 mg, 200 mL glycerol (10% vol.)8600 µmol g−1Bare β-Bi2O3 and g-C3N4
counterparts
>20 [157]
17Pt/CN-A150 composite300 W Xe, 10 mg, 100 mL DI water containing
20 vol% TEA
1150.8 µmol h−1Pt/CN-PR
18.2 µmol h−1
g-C3N4
63.2
4.6
[158]
18g-C3N4@Ni3Se4
gC3N4@CoSe2
5 W LED, 10 mg, 30 mL 15% v/v TEOA16.4 µmol∙h−1
25.6 µmol∙h−1
Pristine g-C3N4
1.9 µmol∙h−1
8
13
[159]
19MoS2/g-C3N4300 W Xe, 5 mg, 40 mL DI water, 10% v/v of TEOA1787 mmol h−1 g−1MoS2
g-C3N4
6
40
[160]
20Carbon vacancies containing g-C3N4300 W Xe, 100 mg, 90 mL
deionized water, 10 mL TEOA, 3 wt% Pt
450 µmol h−1 g−1Pristine g-C3N4 225 µmol h−1 g−12 [161]
213 wt% La2NiO4/g-C3N4300 W Xe, 10 mg,
80 mL 20 vol% methanol
312.8 µmol h−1 g−1La2NiO4 5.8 µmol h−1 g−1
g-C3N4 7.1 µmol h−1 g−1
53.8
43.9
3.7% 420 nm[162]
2218% Ag/AgBr/g-C3N4300 W Xe, 50 mg, 90 mL deionized water, 10 mL TEOA1587.6 µmol h−1 g−1g-C3N4
59.1 µmol h−1 g−1
26.9 [163]
23g-C3N4/N-doped carbon300 W Xe, 10 mg, 20 mL 10 vol% TEOA, 0.5 wt% Pt23.0 µmol h−1g-C3N4/C 5.9 µmol h−14 [164]
24Dendritic fibrous nanosilica/g-C3N4-0.5300 W Xe, 10 mg, 0.019 M, 80 µL K2PtCl4, 5 mL TEOA4662 µmol h−1 g−1Pristine g-C3N47 [165]
25Ag0.1Pd0.9/2D CNNs300 W Xe, 100 mg, 3:1 FA/SF (1.0 M, 4 mL)231.6 mmol h−1Ag0.1Pd0.9/2D CNNs under no light1.8727.8% 400 nm[166]
26g-C3N4/WO3/WS2300 W Xe, 20 mg,
100 mL (20 vol%) TEOA
29 µmol h−1 g−1g-C3N4 nanosheets7.88.9% 420 nm[167]
27CeO2/g-C3N4 -6500 W Xe (400 nm), 0.1g, 100 mL 0.35 M Na2SO3 and 0.25 M Na2S1240.9 µmol h−1 g−1Pure CeO25.2 [168]
28Nitrogen vacancies-g-C3N43 W 420 nm LED, 0.02 g, 90 mL H2O, 10 mL TEOA, 1% H2PtCl6 2H2O (10 mg/mL)3259.1 µmol h−1 g−1Pristine g-C3N48.7 [169]
29Black Cu-g-C3N4
nanosheets composite
300 W Xe, 10 mg
100 mL (1:9 TEOA: Water)
526 µmol h−1 g−1g-C3N4: 280 [170]
30Amino-group-rich porous g-C3N4 nanosheets (AP-CN 1.0)420-nm LED, 0.05 g, 80 mL 10 vol% TEOA, 1 wt% Pt130.7 µmol h−1Bulk g-C3N44.95.58[171]
310.3-MoS2/g-C3N4300 W Xe, 100 mg, 50 mL, deionized water and 5 mL TEOA12 mmol h−1 g−1Pristine g-C3N4
g-C3N4 (Pt)
218
3
0.5% 420 nm[172]
32g-C3N4/ZIF-67MaX 303 solar simulator, 20 mg, 0.5 M Na2SO42084 μmol g−1Bare g-C3N4
541 μmol g−1
3.84 [141]
332D/2D ZnCoMOF/g-C3N4300 W Xe, 10 mg, 0.1 mL DMF1040.1 Bulk g-C3N4: 33.2
2D g-C3N4: 3.5
[116]
34PCN-222(M)/g-C3N4300 W Xe, 10 mg, 25 mL TEOA1725.5 µmol h−1 g−1 PNi: 19.3
CN: 3.7
[115]
Table 2. Photocatalytic CO2 reduction over g-C3N4-based materials.
Table 2. Photocatalytic CO2 reduction over g-C3N4-based materials.
EntryPhotocatalystExperimental
Details
Productivity
/μmol·g−1·h−1
Reference
Material
/μmol·g−1·h−1
Enhancement Relative to
Conventional g-C3N4
Apparent
Quantum
Efficiency/%
Ref.
1Ni/g-C3N4-0.5 catalyst300 W Xe, 94–95 kPa,
10.0 mg, deionized water
CO: 19.9g-C3N4: 4.84.1 [192]
2S-scheme CuWO4 @ g-C3N4 core-shell microspheres300 W Xe (≥420 nm), 0.1 g NaHCO3, 0.5 mL 4M HClCO: 4.15
CH4: 0.12
g-C3N4
CO: 1.56
CH4: 0.02
2.7 [193]
3Hydroxyl-modified g-C3N4/flower-like
Bi2O2CO3 composites
blue LED (4 × 3 W) 450 ± 20 nm, 40 mg, deionized waterCO: 26.69Pristine g-C3N4
CO: 1.47
18.2 [194]
4Z-scheme g-C3N4/BiVO4 (CN/BVO) heterojunction300 W Xenon lamp, 0.05 g, 5 mL waterCO: 48Pristine BVO
CO: 2
24 [195]
5Ultrathin dimension-matched S-scheme Bi3NbO7/g-C3N4 hetero-structureSolar simulator, 50 mg, deionized water, 1.3 g Na2CO3, 2.0 mL H2SO4CH4: 37.59Ultrathin g-C3N4 nanosheets
CH4: 2.5
15 [196]
6Van der Waals (vdW) heterojunction composite combining g-C3N4 with nitrogen vacancies
and Tp-Tta COF
300 W Xe, 20 mg, 15 mg bpy, 1 μmol CoCl2, acetonitrile, water, TEOACO: 11.25Pristine g-C3N4
CO: 0.25,
g-C3N4 (NH)
CO: 3.5
Pristine g-C3N4: 45
g-C3N4 (NH): 3.2
[197]
7C-NHx-rich
24 g-C3N4
300 W Xe (420 nm), 10 mg g-C3N4, 10 mL deionized water, pH at 30 ℃CO: 185.7g-C3N4
CO: 2.5
g-C3N4: 74 [198]
8g-C3N4/3DOM-WO3300 W Xe (≥420), water, 0.1 g catalyst, 2 mL deionized waterCO: 48.7
CH4: 7.5
O2: 44.5
Pure g-C3N4 nanosheets
CO: 25.2
CH4: undetected
Pure g-C3N4 nanosheets
CO: 1.9
[199]
9g-C3N4/rGO composites300 W Xe, 3 mg mL−1 catalysts, 5 mL
0.2 M NaHCO3, illuminated 12 h
CH3OH: 114CdIn2S4/g-C3N4
CH3OH: 42.7
CdIn2S4/g-C3N4: 2.670.63[200]
1015% LaCoO3 loaded g-C3N435 W Xe (420 nm), 50 mg photocatalyst, pressure 0.30 barCO: 135.2
CH4: 48.5
Pristine La-CoO3
CO: 110
CH4: 28.5
g-C3N4
CO: 114
CH4: 30.4
Pristine
LaCoO3
CO: 1.2
CH4: 1.7
g-C3N4
CO: 1.18
CH4: 1.59
[201]
11Bi2O2(NO3)(OH)/g-C3N4300 W Xe, 20 mg samples, 3 mL DI waterCO: 14.84BON
CO: 0.94
g-C3N4
CO: 3.29
pure BON:15 g-C3N4: 3.5 [202]
12Z-scheme SnS2/gC3N4/C300 W Xe, 0.05 g catalyst 100 mL deionized water, 25 °C, 5 hCO: 40.86Pristine g-C3N4
CO: 7.42
Pristine g-C3N4
5.5
[203]
13ND/g-C3N4300 W Xe (>420 nm), 30 mg catalyst, 18 mL acetonitrile, 6 mL water, 1 μmoL
CoCl2⋅6H2O
CO: 10.98CO: 0.59bulk g-C3N4
18.6
[204]
14ZnIn2S4 nanosheets modified hexagonal g-C3N4 tubes300 W Xe (420 nm), 4 mg, 2 mL water, 1 mL of triethanolamine, 3 mL acetonitrile, 15 mg 2′2-bipyridine (bpy) and 2 µmol of CoCl2CO: 883HCNT: 66
ZIS: 367.9
HCNT: 13
ZIS: 2.4
8.9%[205]
15g-C3N4/covalent triazine framework (CN/CTF 2.5%)300 W Xe, 5 mg catalyst in 4 mL acetonitrile, 1 mL Co(bpy)3Cl2 triethanolamineCO: 151.1CTF: 5.93
CN: 60.44
CTF: 25.5
CN: 2.5
[206]
16g-C3N4-W18O49 nanocomposite300 W Xe, 50 mg catalyst in 1 mL deionized waterCH4: 1.38g-C3N4: 0.17 W18O49: 0.12g-C3N4: 8.12
W18O49: 11.5
[207]
17SnS2/Au/g-C3N4 embedded structure300 W Xe, 20 mg, 100 mL water and TEOA, 140 kPaCO 93.81
CH4 74.98
[208]
18Bi3O4Cl/20%g-C3N4300 W Xe, 0.05 g catalyst, 5 mL H2OCO: 6.6
CH4: 1.9
Pure g-C3N4 CO: 2.2,
CH4: 0.6 Bi3O4Cl
CO: 2.9
CH4: 0.7
g-C3N4
CO: 3
CH4: 3.17
Bi3O4Cl
CO: 2.28
CH4: 2.71
Bi3O4Cl/20% g-C3N4 is 0.14% under 365[209]
192D/2D g-C3N4/NaBiO32H2O (10 CN/NBO)300 W Xe, 25 mg, deionized water, 1.2 g
NaHCO3, 2 mL H2SO4
(1:1 vol)
CO: 110.2
CH4: 43.8
Pure CN
CO: 65.68
CH4: 0.42
NBO
CO: 26.45
CH4: 4.81
Pure CN
CO: 1.68
CH4: 104.3
NBO
CO: 4.16
CH4: 9.1
[210]
20Ultrathin nanosheet g-C3N4 (NS-g-C3N4)300 W Xe (420 nm), 0.1 g photocatalyst, 50 mL 50 g/L
KHCO3
CO: 38 μmol/L with 6 hBulk g-C3N4
CO: 6.56 μmol/L
CO: 5.8 [211]
213% CdS-g-C3N4 heterostructures300 W Xe (420 nm), 1 g/L catalyst 100 mL H2O, 80 ℃, 125 mg Na2CO3, 0.25 mL HCl (4 M)CH3OH: 192.7CdS
CH3OH: 47.1
pristine
g-C3N4
CH3OH: 32.6
CdS: 4.1
pristine
g-C3N4: 5.9
[212]
22Z-scheme ZnO/Au/g-C3N4 micro-needles film (3-ZAC)300 W UV–vis lamp, fiberglass sheets, 0.4 M Pa86.2 μmol m−2 h−1Pure ZnO
19.16 μmol m−2 h−1
Pure ZnO film: 4.5 [213]
23rGO/R-CeO2/g-C3N4300 W Xe, 100 mg catalysts, 100 mL 1 M NaOH, 1 mmol TEOA, 0.4 MPaCO:15.8
CH4: 8.15
CO: 3.95
CH4: 1.36
Pure g-C3N4
CO: 4
CH4: 6
[214]
24g-C3N4/ZnO composites300 W Xe (λ ≥ 420 nm), 60 mg catalysts, 1.60 g NaHCO3, H2SO4 (40%, 5.0 mL)CH4: 19.8
CO: 0.37
g-C3N4
CH4: 0.9
CO: 4.8
g-C3N4
CH4: 22
CO: 0.078
[215]
25K-CN-7300 W Xe, 50 mg catalyst, 200 μL deionized water, 1 cm × 3 cm ITO glass;
0.5 M Na2SO4
CO: 8.7Ordinary g-C3N4
CO: 0.348
Ordinary g-C3N4: 25 [216]
26g-C3N4/CdS
heterostructure nanocomposite
150 W Xe, 20 mg catalyst,
7 mL acetone nitrile, 0.5 mL H2O, 0.5 g TEOA, 4 µmol [Co(bpy)3]Cl
CO: 234.6CN-12: 58.65
CdS: 9.2
CN-12: 4.0
CdS: 25.5
[217]
27Porous structure g-C3N4 with nitrogen defect photocatalysts (DCN-P)300 W Xe, 0.05 g catalyst, 100 mL deionized waterCO, 19.7
CH4: 37.1
Bulk g-C3N4
CO: 4.1
CH4: 9.6
Bulk gC3N4
CO: 4.8
CH4: 3.86
[218]
28g-C3N4/Bi2O2[BO2(OH)]
(CNBB-3)
300 W Xe, 20 mg sample, 2 mL deionized water, 1.7 g Na2CO3, 15 mL H2SO4CO: 6.09Pristine g-C3N4
CO: 2.19
Pristine g-C3N4
2.78
[219]
29Type-Ⅱ heterojunction of Zn0.2Cd0.8S/g-C3N4300 W Xe, 80 °C, 0.6 MPa, 10 mg catalyst, 20 mL H2OCH3OH: 11.5 ± 0.3Zn0.2Cd0.8S: CH3OH: 4.4 ± 0.2
g-C3N4:
CH3OH: 4.2 ± 0.1
Zn0.2 Cd0.8S: 2.6
g-C3N4: 2.7
[220]
303ZIF/1.5Au-PCN300 W Xe, 0.1 g, 50 mL H2OCO: >10
CH4: >4
Pristine g-C3N4
8
[221]
31TPVT-MOFs@g-C3N4-10LED light, 1 mg, 1 mL dichloromethaneCO: 56.4Pure g-C3N4: 17.5Pure g-C3N4
3.2
[114]
32NH2-MIL-101(Fe)/g-C3N4-30 wt%300 W Xe, 2 mgCO: 132.8g-C3N4:
19.2
g-C3N4
6.9
[222]
Table 3. Photocatalytic degradation of pollutants over g-C3N4-based materials reported within the last three years.
Table 3. Photocatalytic degradation of pollutants over g-C3N4-based materials reported within the last three years.
EntryPhotocatalystPollutant ConcentrationLight SourceDegradation
Efficiency/%
Ref.
15% g-C3N4-TiO2Acetaminophen: 0.033 mM300 W Xe (>400 nm)99.3 in 30 min[233]
23ZIF/1.5Au-PCNBisphenol A350 W Xe (>420 nm)>85%[221]
3Cu(tmpa)/20%CNCongo red: 100 mg·L−1150 W Xe98.2% in 3 min[234]
4BiO-Ag(0)/C3N4@
ZIF-67
Congo red: 12 mg·L−1Natural sunlight90% in 150 min[13]
5C3N4/RGO/Bi2Fe4O9Congo red: 10 mg·L−1LED 30 W87.65% in 60 min[235]
6g-C3N4/Co-MOFCrystal violet: 4 ppmMaX 303 solar simulator (50 mW/cm)95% in 80 min[141]
7Honeycomb-like
g-C3N4/CeO2-x
Cr (VI): 20 mg·L−1300 W Xe (>420 nm)98%
in 150 min
[236]
8Sm6WO12/g-C3N4Levofloxacin: 10 mg·L−1150 Mw cm−2 tungsten lamp98% in 70 min[237]
9O-g/C3N4Lincomycin: 100 mg·L−1PCX50C system (>420 nm)99% within 3 h[238]
10ZnO-modified g-C3N4Methylene blue: 10 ppm200 W tungsten
lamp (>420 nm)
97% in 80 min[239]
11Wood-like g-C3N4@WDCMethylene blue: 20 mg·L−1300 W Xe (>400 nm)98% in 60 min[240]
12BiO-Ag(0)/C3N4@
ZIF-67
Methylene blue: 12 mg·L−1Natural sunlight96.5% in 120 min[13]
13Cerium-based GO/g-C3N4/Fe2O3Methylene blue: 10 mg·L−1Light bulb70.61% in 45 min[14]
14Ytterbium oxide-based GO/g-C3N4/Fe2O3Methylene blue: 10 mg·L−1Light bulb83.5% in 45 min[14]
15Cu(tmpa)/20%CNMethylene blue: 10 mg·L−1150W Xe92.0% within 20 min[234]
16C3N4x/AgOy@Co1-xBi1-yO7Methylene blue: 25 mL 10 mM100 W tungsten bulb96.4% in 120 min[12]
17Ternary composites of Zr-MOF combined with g-C3N4 and Ag3PO4Methylene blue: 10 mg·L−185-watt tungsten lamp
outdoor/solar light in
an open air
95% within 240 93% within 105 min [241]
18PSCN/Ag@AgI/WO3Malachite green: 1 × 10 −4 mol dm−335 W LED90% in 60 min[242]
19Cu(tmpa)/20%CNMalachite green: 30 mg·L−1150W Xe92.9% in 35 min[234]
2020% g-C3N4/Bi4O5I2Methyl orange: 20 mg·L−1350 W Xe0.164 min−1[243]
21Cu(tmpa)/20%CNMethyl violet: 10 mg·L−1150W Xe92.0% in 60 min[234]
22MnCo2O4/g-C3N4Nitrobenzene: 40 mg L−1CMCN2/PMS system96.7% in 240 min[244]
23C3N4x/AgOy@Co1-xBi1-yO7Oxytetracycline: 25 mL 25 mM100 W tungsten bulb93% in 160 min[12]
24g-C3N4/WO3/WS2Rhodamine B: 25 mg L−1300 W Xe (>420 nm)96.2% in 20 min[167]
25Flower-like Bi12TiO20/g-C3N4Rhodamine B: 20 mg·L−1150 mW·cm−2 Xe (>420 nm)100% in 30 min[245]
26CdS/CQDs/g-C3N4Rhodamine B: 10 mg·L−1300 W Xe
(>420 nm)
100% in 20 min[246]
27Ytterbium oxide-based GO/g-C3N4/Fe2O3Rhodamine B: 10 mg·L−1Light bulb67.11% in 45 min[14]
28Cerium-based GO/g-C3N4/Fe2O3Rhodamine B: 10 mg·L−1Light bulb63.08% in 45 min[14]
29Fish-scale g-C3N4/ZnIn2S4Tetracycline: 10 mg·L−1300 W Xe (>420 nm)74% in 30 min[247]
31Flower-like Co3O4/g-C3N4Tetracycline: 15 mg·L−1350 W Xe (>420 nm)85.32% in 120 min[248]
3110 wt% CuAl2O4/g-C3N4Tetracycline hydrochloride: 100 mg·L−1300 W Xe (>400 nm)89.6% in 60 min[249]
32CO-C3N4Tetracycline hydrochloride: 10 mg·L−1300 W Xe (>420 nm)97.77% (PMS) in 40 min[250]
33ZIF-67/g-C3N4Venlafaxine: 10 mg·L−1-27.75% within 120 min[251]
34ZIF-67/MIL-100(Fe)/g-C3N4Venlafaxine: 10 mg·L−1-100% within 120 min[251]
35ZIF-67/MOF-74(Ni)/g-C3N4Venlafaxine: 10 mg·L−1-91.8% within 120 min[251]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, Q.; Li, Y.; Huang, F.; Song, S.; Ai, G.; Xin, X.; Zhao, B.; Zheng, Y.; Zhang, Z. Recent Advances in g-C3N4-Based Materials and Their Application in Energy and Environmental Sustainability. Molecules 2023, 28, 432. https://doi.org/10.3390/molecules28010432

AMA Style

Wang Q, Li Y, Huang F, Song S, Ai G, Xin X, Zhao B, Zheng Y, Zhang Z. Recent Advances in g-C3N4-Based Materials and Their Application in Energy and Environmental Sustainability. Molecules. 2023; 28(1):432. https://doi.org/10.3390/molecules28010432

Chicago/Turabian Style

Wang, Qian, Yongfei Li, Fenglin Huang, Shaofu Song, Ganggang Ai, Xin Xin, Bin Zhao, Yajun Zheng, and Zhiping Zhang. 2023. "Recent Advances in g-C3N4-Based Materials and Their Application in Energy and Environmental Sustainability" Molecules 28, no. 1: 432. https://doi.org/10.3390/molecules28010432

Article Metrics

Back to TopTop