Next Article in Journal
New QSPR/QSAR Models for Organic and Inorganic Compounds: Similarity and Dissimilarity
Previous Article in Journal
Synthesis of Physically Activated Carbons from Vitellaria paradoxa Shells for Supercapacitor Electrode Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modification Strategies of g-C3N4-Based Materials for Enhanced Photoelectrocatalytic Degradation of Pollutants: A Review

1
Key Laboratory of New Processing Technology for Nonferrous Metal & Materials, Ministry of Education, Guangxi Key Laboratory of Optical and Electronic Materials and Devices, College of Materials Science and Engineering, Guilin University of Technology, Guilin 541004, China
2
College of Environmental Science and Engineering, Guilin University of Technology, Guilin 541004, China
*
Authors to whom correspondence should be addressed.
Inorganics 2025, 13(7), 225; https://doi.org/10.3390/inorganics13070225
Submission received: 4 May 2025 / Revised: 10 June 2025 / Accepted: 1 July 2025 / Published: 3 July 2025

Abstract

Graphite carbon nitride (g-C3N4) is a low band gap non-metallic polymer semiconductor that has broad application prospects and is an ideal material for absorbing visible light, as g-C3N4 materials have strong oxidation properties and are easy to modify. The structure formation of g-C3N4-based materials makes a series of photocatalytic synthesis reactions possible and improves photocatalytic reaction activity. In this paper, the development history, structures, and performance of g-C3N4 are briefly introduced, and the modification strategies of g-C3N4 are summarized to improve its photocatalytic and photoelectric catalytic properties via doping, heterojunction construction, etc. The light absorption and utilization of the catalysts are also analyzed in terms of light source conditions, and the application of g-C3N4 and its modified materials in photocatalysis and photocatalytic degradation is reviewed.

Graphical Abstract

1. Introduction

At present, pollutants caused by rising energy use and human activity are currently one of the most serious environmental issues. The world is trying to use solar energy for energy conversion, environmental protection, and water purification [1]. Various environmental pollutants (dyes [2], pesticides [3], drugs [4], phenols [5], metal ions [6], etc.) in water are toxic and recalcitrant in the environment, and their removal is of great concern [7].
Semiconductor photocatalytic technology is precisely based on the chemical conversion and storage of solar energy as the core. Removing toxic and harmful pollutants through photocatalysis will provide people with a green and environmentally friendly living space. In contrast, the conversion of solar energy into hydrogen as well as other chemical energy sources, such as photolysis of water and carbon dioxide conversion, is expected to completely solve the energy crisis brought about by the increasing depletion of energy. However, the low utilization efficiency of absorbed light and the high carrier recombination seriously restrict the practical application of photocatalytic technology. Photocatalysis technology is the foundation of photoelectrocatalysis (PEC), which can effectively promote the separation of photogenerated electron-hole pairs by utilizing photoelectric synergy, which has the advantages of easy separation and recovery. According to studies, applying a certain bias voltage to semiconductor materials can segregate the photogenerated charge while significantly extending the life of the photogenerated carrier, and the photocatalytic efficiency will thereafter be significantly increased [8,9,10]. Currently, photoelectrocatalytic technology is widely used in the fields of multi-pollutant degradation, carbon dioxide reduction, water oxidation, and hydrogen production [11,12,13].
Carbon materials (graphene, g-C3N4, carbon dots, etc.) have excellent electrical conductivity [14,15,16]. Through doping elements, efect engineering, grafting functional groups, and other methods, they can easily design their surface chemical properties and regulate their structure. The interface interaction between the inorganic carbon material and semiconductor can enhance the semiconductor’s photo-response performance, photogenerated carrier migration characteristics, adsorption efficiency, and catalytic active site, all of which contribute to the semiconductor’s photoelectrocatalytic performance [17]. Because of this, carbon materials have a promising future in photoelectrocatalytic degradation of pollutants and other areas [18]. As g-C3N4 is a unique 2D layered nonmetallic material [19,20,21], it offers the benefits of a straightforward synthesis process and strong thermal stability, and its energy band structure is highly suitable for the two key steps of hydrogen and oxygen production in the photocatalytic decomposition of water. As a photocatalytic material, g-C3N4 has broad application prospects [22,23,24]. As shown in Figure 1, g-C3N4 exhibits significant application potential in the degradation of both water and gas pollutants. In terms of water pollutant degradation, it demonstrates good treatment effects for both organic and inorganic contaminants. For organic pollutants, g-C3N4 can oxidatively decompose dye molecules such as methylene blue (MB), rhodamine B (RhB), and methyl orange (MO) in dye wastewater under light irradiation, while for antibiotic and pharmaceutical residues like tetracycline (TC), ciprofloxacin (CIP), and ibuprofen, it can mineralize them into CO2 and H2O, avoiding secondary pollution in traditional water treatment. For phenolic compounds and persistent organic pollutants (POPs) such as phenol, bisphenol A (BPA), and p-nitrophenol (4-NP), some intermediate products remain toxic, necessitating catalyst optimization to achieve complete mineralization. For inorganic pollutants, g-C3N4 can utilize photogenerated electrons to reduce heavy metal ions, such as by converting highly toxic Cr(VI) to Cr(III). In gas pollutant degradation, nitrogen oxides (NOx) and sulfides (H2S) such as NO, SO2, and H2S can be degraded through photocatalytic oxidation (e.g., NO→NO3) and surface adsorption activation to promote gas molecular decomposition. For volatile organic compounds (VOCs) in indoor and outdoor air purification, constructing porous g-C3N4 to increase gas adsorption sites and compositing with metal oxides to enhance oxidation capacity are effective improvement strategies.

2. Development History and Structures of Graphite Carbon Nitride

C3N4, a polymer derivative containing carbon and nitrogen (C6N9H3), is considered to be one of the first chemicals synthesized, and it was reported in 1830 [25]. Jöns Jakob Berzelius heated the “mystery” mixture (potassium thiocyanate with a large amount of AlCl3), and when the temperature was about 350 °C, a molten product was produced which could be successfully isolated with KOH [26,27]. Justus von Liebig then suggested naming it “melon”, which objectively made it one of the oldest artificial polymerized derivatives [28]. Franklin first proposed the concept of “carbon nitride” in 1922 when he obtained an amorphous carbon-nitrogen compound by pyrolyzing precursors such as Hg(CN)2 and Hg(SCN)2 [29]. In 1937, L. Pauling and J. H. Sturdivant first proposed that C3N4 is a cluster compound with coplanar triazine as the basic structural unit, and this view was proven through X-ray crystallography [30]. In the 1990s, C3N4 came back into the focus of developers after a long period of neglect, but back then, it was all about the breakthrough of the superhard material β-C3N4 [31,32]. In 1985, A.Y. Liu and M.L. Cohen [33] had theoretically predicted the structure of β-C3N4 by replacing Si with C atoms, based on the existing crystal structure of β-Si3N4. They deduced that the material had a hardness comparable to diamond. In 1996, D. M. Teter and R. Hemley recalculated the structure of C3N4 using first principles and proposed that C3N4 has five structures [34], namely α phase (P31c), β phase (P3), c phase (I 4 ¯ 3d), p phase (P 4 ¯ 2m), and g phase (P 6 ¯ m2). Among these, the first four are superhard materials, while g-C3N4 is not. Compared with traditional metal catalysts, g-C3N4 has the advantages of high stability, acid and alkali resistance, and convenient modification, making it extremely suitable for inorganic metal catalyst design and synthesis [35,36]. In 2006, g-C3N4 began to be used in the field of heterogeneous catalysis. In 2009, the research group of Professor X.C. Wang [37] from Fuzhou University proved that g-C3N4 non-metallic semiconductors can achieve photocatalytic hydrogen production. In 2013, J. Xu et al. [38] prepared g-C3N4 nanosheets with a single atomic layer structure via the chemical exfoliation method, which exhibit stronger intrinsic and novel properties and thus shows great potential for application in many fields. The photocatalytic H2 production activity and pollutant decomposition activity of monolayer g-C3N4 nanosheets are much higher than that of bulk g-C3N4, which has great potential for photocatalysis and photosynthesis [39,40].
g-C3N4 is in the form of powder with a yellowish surface. This material has a low density and a thermal decomposition temperature of 600 °C, and it is incompatible with water. Under light conditions, g-C3N4 nanomaterials are mixed with the water phase to form hydroxyl radicals, which have strong oxidizing properties and can degrade organic pollutants at the same time. In recent years, thanks to the efforts of researchers, the practical application of g-C3N4 in the field of photocatalysis has been expanded, and even mass production has been successfully achieved [41].
Among the five structural compositions of carbon nitride, g-C3N4 is the easiest material for achieving practical applications due to its band gap width and adjustable band structure. Its band gap is 2.7 eV, which is much smaller than those of other types. g-C3N4 has a two-dimensional layered structure similar to graphite stacks, and it was shown that two different polymer structures can be used as the basic structural units of each lamella, being connected by terminal N atoms to form an infinitely extended plane. Currently, there are two types of g-C3N4 nanomaterials, S-triazine units and 3-s-triazine units, in which the g-C3N4 nanomaterials based on 3-s-triazine units are much more stable, and this result has been confirmed by Jacob et al. [42], who investigated the calculation of the electronic band gap for both g-C3N4 and the triple-s-triazine-based g-C3N4 by using the multibody Green’s function method.
But it has been shown both experimentally and theoretically that melon, i.e., the polymorph with partial condensation of heptazine monomers, is the structure most frequently synthesized through thermal treatments [43,44]. B.V. Lotsch et al. [45], on the other hand, suggested that this model is inaccurate, and they argued that melons are planar layers and compounds consisting of carbon (IV) nitride cores (heptazine units) are regarded as “defective g-C3N4” materials. Since the preparation of g-C3N4 from hydrogen-containing precursors essentially yields materials with similar structural features and hydrogen contents, the graphite carbon nitride synthesis products should actually be polymers such as melons or related compounds. The discussion on g-C3N4 has a focus of many researchers. g-C3N4 is one of the representative nonmetallic semiconductors [46] which have received much attention this decade for its application in photocatalytic environmental remediation.

3. Graphitic Carbon Nitride Modification

Although g-C3N4 has the advantages of a narrow band gap, low cost, and non-toxicity, pure g-C3N4 still has the disadvantages of a narrow absorption spectral range and a high photogenerated electron-hole recombination rate [47], which impede its application in engineering practice. Compared with other means, the modification process of g-C3N4 via elemental doping greatly simplifies and streamlines the experimental steps and is more effective in broadening its light absorbance efficiency [48]. The formation of heterojunctions between g-C3N4 and one or more different substances does not only inhibit the complexation of photogenerated carriers; the synergistic effect between g-C3N4 and other components in the heterojunctions can give the photocatalysts some new characteristics [49]. Therefore, this section will focus on outlining three modifications of g-C3N4: doping and constructing heterojunctions, as shown in Figure 2. Table 1 shows a comparison of them.

3.1. Dopant Modification

3.1.1. Metal Doping

Metal ion-doped g-C3N4 inhibits electron-hole recombination, improves its optical and physicochemical properties, extends the absorption wavelength range to the visible range, and also increases its specific surface area [50]. N atoms with a lone pair of electrons in the g-C3N4 triazine structure can easily form coordination bonds with metal cations and improve the performance of photocatalysts by lowering the band gap of the material [51]. Doped metals include alkali metals, transition metals, and rare metals. Metal ions are mainly located in the vacant region of g-C3N4 because g-C3N4 has a larger atomic radius than the atomic radii of nitrogen and carbon in the structure [52,53]. Cavity doping occurs when metal ions enter the triangular pores between g-C3N4-linked triazine structures, leading to strong coordinative interactions between metal ions and negatively charged nitrogen atoms. Metal ion-doped g-C3N4 nanocomposites can be used for the rapid degradation of organic pollutants in wastewater [54]. Figure 3 presents the optimized structures of intrinsic g-C3N4 and doped alkali metal g-C3N4, with some metal ions as examples (Li, Na, K, and Rb).
H. Zhang et al. [55] prepared metal-doped g-C3N4 through the solvothermal method, and S. Ye et al. [56] successfully synthesized photocatalytic Fe2O3/g-C3N4 composites. They demonstrated that K-doped g-C3N4, Rb-doped g-C3N4, and Fe2O3/g-C3N4 have lower forbidden bandwidths and strong light-absorbing properties compared with pure g-C3N4. Y.Z. Zhang et al. [57] prepared sodium-doped graphite nitride. It was found that sodium doping significantly improved the photocatalytic ability of H2, which was 9.2 times higher than that of graphite nitride. The experiments of P.Q. Deng et al. [58] realized doped Ni2+ in the cavities of the g-C3N4 framework. The increase in Ni content affects the possible gap sites of the doped Ni and changes the positions of its valence band (VB) and conduction band (CB). The Cu+- g-C3N4 complex prepared by J.Q. Ma et al. [59] has high Fenton activity, and its mechanism is shown in Figure 4. Cu2+ is reduced to Cu+ by a hydroxyl radical mechanism and finally embedded in the lamellar g-C3N4 via coordination with pyridine N. The bonded Cu+ participates in the generation of 1O2 and HO·, and then, due to its structural stability, it can be rapidly reduced to the cuprous state and thus has a high catalytic ability in Fenton-like reactions.
L.W. Ruan [60] computationally investigated the optical properties of lithium-doped g-C3N4 monolayers. There is no LUMO or HOMO in lithium atoms. Therefore, the active sites will not change in the doped systems compared to the pristine g-C3N4. However, by calculating LUMO and HOMO, it was found that doped Li atoms can enhance the hybridization of LUMO and HOMO values, indicating that the light-generated electron and hole pairs can be effectively separated. Ta et al. [61] synthesized metal-doped ZnO/g-C3N4 composites with a sponge-like porous structure via a facile one-pot pyrolysis method, among which the magnesium (Mg)-doped composite exhibited the highest photocatalytic performance. A study by Li et al. [62] demonstrated that Ag-doped g-C3N4-TiO2 exhibits superior photocatalytic degradation performance due to its narrower band gap compared with pure g-C3N4-TiO2, as well as faster separation and transfer of photogenerated charge carriers. Ji et al. [63] experimentally derived the photo-Fenton reaction mechanism of Fe-doped g-C3N4 as follows. Firstly, Fe-g-C3N4 generates photo-induced electron-hole pairs under light irradiation. The photo-induced holes of Fe-g-C3N4 can oxidize H2O/OH to generate ·OH. Fe doping remarkably improved the redox properties. Owing to the interfacial charge transfer effect, the Fe3+ is reduced into Fe2+ by photo-excited electrons, followed by the reaction of Fe2+ with H2O2, which forms the ·OH radical and effectively oxidizes RhB. Doping Fe decreases the photoelectron-hole complex efficiency, and the transfer of photoelectrons to Fe improves the photocatalytic activity of g-C3N4. As shown in Table 2, metal doping modification can significantly enhance the photocatalytic degradation performance of g-C3N4, and multiple metal-doped systems exhibit higher degradation rates in pollutant degradation.
Modification of g-C3N4 using metal doping can enhance the catalyst degradation rate. However, it has been found that neither mono nor bimetallic metals can change the g-C3N4 structure, and relying only on gap trapping of electrons and inhibition of some of the photo complex reactions cannot completely and drastically enhance the catalytic efficiency. Therefore, an attempt can be made to fully merge the respective advantages of metal and g-C3N4 into one to explore the doping and application studies of multi-metals.

3.1.2. Non-Metal Doping

Non-metallic atoms have high ionization energies and different electronegativities. When introduced, non-metallic atoms accept electrons from other substances to form covalent bonds, thus changing the electron distribution around the dopant sites [64,65]. In general, highly electronegative dopant atoms promote the migration of electrons from neighboring C atoms to the dopant sites; the original chemical inertia is destroyed, and the C atoms will have induced polarization to form new reaction sites. When the electronegativity of the dopant atoms is low, it can increase the asymmetric spin density of the neighboring C atoms, thus increasing the electron-donating ability of g-C3N4 [66,67]. Thus, non-metallic doping (O, N, P, C, S, B, etc.) enables reactions with other species to form covalent bonds, thereby altering electron distribution around dopant sites via charge polarization [68]. The photogenerated electron mobility of the material is increased, and the electron-hole complexation rate is decreased, which effectively adjusts the electronic and molecular structure of g-C3N4, expands its light responsive range [69], and ultimately improves the photocatalytic properties.
In M. Arumugamm’s research [70], g-C3N4 doped with nonmetals (B, O, P, and S) was prepared, and various physicochemical analyses were performed, which confirmed that the doped nonmetallic elements could improve the performance of photocatalysts by improving charge separation. Its efficiency in reducing carbon dioxide and water to CH4 was 1~2 times that of the original. The S-g-C3N4 prepared by G. Liu et al. [71] was also effective in catalyzed phenol degradation at a light source with a wavelength > 400 nm. J. Marie and D.W. Dikdim et al. [72] investigated activated carbon/g-C3N4 and found that the addition of activated carbon effectively inhibited the electron-hole pair complexation. A non-metallic photocatalyst was prepared by J.J. Zhang et al. [73]. The experimental results showed that PI-g-C3N4 (perylene imide-modified g-C3N4) could realize the activation of persulfate to degrade BPA under visible light. The above studies have shown that doping of most of the nonmetals (e.g., S, P, O, C, I, and B) can effectively narrow the band gap of g-C3N4 and thus broaden the optical absorption spectral range of the material, making it easier to be excited by visible light [74]. However, when the band gap of the material is too small, the redox potential of the valence or conductance band of the catalytic material decreases [75], which is also detrimental to the photocatalytic property of the material. Therefore, the ideal outcome of g-C3N4 band gap regulation should be that the regulated material has an appropriate band gap (~2.0 eV) [76] which can fully absorb sunlight while having appropriate conduction and valence band positions such that the photogenerated carrier has a sufficient redox capacity and the synergy of different properties can produce better effects. Figure 5 shows the band gap variation in g-C3N4 and the energy band structure of g-C3N4 with different nonmetal doping.
Non-metal doping strategies are generally favored due to their low cost and ease of preparation. Although some results have been achieved in the research of non-metallic doping modification of g-C3N4, the current attention on the non-metallic doping modification strategy focuses on single non-metallic doping, and there is less exploration of multi-non-metallic co-doping modification of g-C3N4. Studies on the synergistic promotion of g-C3N4 photocatalysis via non-metallic doping and other modification methods are even less common.

3.2. Constructing Heterojunction Structures

The heterojunction structure of g-C3N4 was constructed to effectively inhibit the carrier composite. A heterojunction is generally defined as the interface between two regions of different semiconductors with unequal band structures and creating interfacial band alignments [78,79,80,81]. The formation of g-C3N4-based heterojunctions often includes the following categories: (1) Type I heterojunction, (2) Type II heterojunction, (3) p-n junction, (4) Schottky junction, and (5) Z-scheme heterojunction. These categories’ energy band structures are shown below in Figure 6 [82]. Based on the understanding of a traditional heterojunction, a new concept of S-type heterojunction has been proposed. The concept was first proposed by Prof. Jianguo Yu’s group in 2019, and its full name is the step-scheme heterojunction, shown in Figure 7 [83,84]. Table 3 presents examples of constructing different heterojunctions and their applications, providing a practical reference for understanding heterojunction - based design strategies.
Table 3. Examples of constructing different heterojunctions and their applications.
Table 3. Examples of constructing different heterojunctions and their applications.
ClassificationTypical ExampleApplication
Type I heterojunctionZnIn2S4/ultrathin-g-C3N4 [85]Photocatalytic H2 production reaction under visible light
CdIn2S4/g-C3N4 [86]Photodegradation of RB19 under visible-light irradiation
Type II heterojunctiong-C3N4/g-C3N4 [87]
(with different raw materials)
Enhancement of photocatalytic H2 production and CO2 reduction activity
g-C3N4@Cs2AgBiBr6 (CABB) [88]CO2 photoreduction process
p-n junctionCo3O4/n-type g-C3N4 [89]CO2 reduction
P-group intercalated g-C3N4 (NP–CN)/Bi2WO6 (BWO), NPB [90]Photodegradation of p-nitrophenol into harmless products
Schottky junctionAg-decorated P-doped g-C3N4 nanosheets (Ag-(P/CNNS)) [91]Water splitting and degradation of rhodamine B (RhB)
g-C3N4/Ti3C2 [92]Efficient visible-light photocatalytic hydrogen
Z-scheme heterojunctionDirect Z-schemeBiOI/g-C3N4 [93]Photocatalytic degradation of phenol
Indirect Z-schemeTiO2/BC/g-C3N4 [94]Photocatalytic reduction of Cr(VI) in aqueous solution
S-scheme heterojunctionC–O-bridged CeO2/g-C3N4 (cCN) [95]Photofixation of N2, H2 energy generation, and methyl orange photodegradation

3.2.1. Type I Heterojunction

In the Type I heterojunction structure (Figure 6a), the energy bands of two semiconductors exhibit a straddling alignment. Electrons migrate to the semiconductor with the lower conduction band edge, while holes migrate to the semiconductor with the higher valence band edge. As an example, K.F. Bai [96] investigated the properties of GaS/g-C3N4 van der Waals (vdWs) heterostructures using density functional theory. The formation energies of GaS/g-C3N4 vdWs heterostructures with different structural configurations were compared, and it was found that the stabilized GaS/g-C3N4 vdWs heterostructures have an I-type band alignment. It was found that the H2 yield of the ternary catalyst NiS/ZnIn2S4/ultrathin-g-C3N4 (NiS/ZIS/UCN) was more than five times higher than that of ZnIn2S4 and that NiS/ZIS/UCN constructed channels with dual high-speed charge transfer to promote photocatalytic H2 generation [95]. As shown in Figure 8, the photocatalytic mechanism, the formation of a Type I heterojunction between ZIS and UCN, and the formation of high-speed charge transfer nano-channels greatly accelerated charge transfer and significantly improved the separation efficiency.

3.2.2. Type II Heterojunction

g-C3N4 and another semiconductor construct a conventional g-C3N4-based Type II heterojunction structure (Figure 6b), in which both the valence and conduction bands of g-C3N4 are above or below the position of the other half of the conductor [97]. The potential difference between the two semiconductors can drive the migration of photogenerated electrons and holes and promote effective space charge redistribution at the interface, which greatly facilitates charge separation and thus greatly improves photocatalytic performance. F. Dong et al. [98] constructed a g-C3N4/g-C3N4 heterojunction composite with different raw materials, a typical type II heterojunction composite structure, and experimentally demonstrated that its visible-light properties were significantly enhanced.

3.2.3. The p-n Junction

The construction of p-n junction structures (Figure 6c) is an effective strategy for the development of noble metal-free hybridized photocatalysts. In the past decades, many novel p-n heterojunctions have been proposed and have shown excellent photocatalytic performance in the oxidative degradation of organic pollutants [99]. According to the theory of semiconductor physics, when p-type and n-type semiconductors with different Fermi energy levels are in contact, an internal electric field is generated at the interface of the p-n junction formed by the two, which has been widely demonstrated to be favorable for the separation of photogenerated carriers [100]. For example, 2D/2D BiOBr/g-C3N4 heterojunctions prepared using self-assembly methods were characterized by p-n junctions [101]. Compared with g-C3N4 alone, the kinetic constants for the degradation of bisphenol A and norfloxacin through the optimized BiOBr/g-C3N4 were improved by factors of 14.74 and 4.01, respectively. Compared with BiOBr alone, the kinetic constants for the degradation of bisphenol A and norfloxacin by the optimized BiOBr/g-C3N4 were improved by factors of 2.20 and 1.36, respectively. The efficient spatial separation and transport of photoexcited electron-hole pairs are important reasons for the good photocatalytic activity and stability of BiOBr/g-C3N4. Hydrothermal pretreatment and the calcination process developed a heterojunction of Co3O4 nanoparticles supported by g-C3N4 microtubules [89]. The p-type Co3O4 and n-type g-C3N4 constructed a p-n junction. The generated internal electric field accelerates the charge transfer. The charge transfer path on the TCN/Co3O4 NP is shown in Figure 9.

3.2.4. Schottky Junction

The Schottky junction catalyst is a heterogeneous structure formed by metal and semiconductors (Figure 6d). Its more obvious characteristic is its rectification properties, which can be adjusted according to the current pattern. Schottky conducted a more systematic study of metal-semiconductor contacts in 1874. He pointed out the existence of a potential barrier on the semiconductor side close to the interface and analyzed it [102,103,104]. The Schottky barrier is able to control the migration of free charge, effectively separating the photogenerated carriers and favoring the participation of holes in the reaction, thus improving the photocatalytic activity. Taking g-C3N4/Ti3C2 as an example, J. Li et al. [92] constructed a kind of unique 2D/3D structure of g-C3N4/Ti3C2, and its photocatalytic hydrogen precipitation mechanism is shown in Figure 10. Due to the presence of a Schottky junction and a wide contact interface in the composite, the excited electrons in the conduction band of g-C3N4 can easily migrate to Ti3C2 under the visible light condition under the action of an internal electric field, while the corresponding holes remain in the valence band of g-C3N4. Subsequently, the higher conductivity of Ti3C2 accelerates the consumption of the trapped electrons.

3.2.5. Z-Scheme Heterojunction

The Z-scheme heterojunction was proposed based on artificial photosynthesis that mimics natural plant photosynthesis. The Z-type system comprises an oxidized photocatalyst and a reduced photocatalyst [105]. When exposed to visible light, electrons on the oxidized photocatalyst CB are transferred and then recombined with holes on the reduced photocatalyst VB, leaving more electrons in the reduced photocatalyst and more holes in the oxidized photocatalyst to participate in the reduction and oxidation reactions [106]. It has the advantages of effectively separating electron-hole pairs, reducing the chance of compounding, retaining more active sites, and expanding the photo-response range. The photocatalytic activity is improved relative to the single photocatalyst of traditional heterojunction photocatalysts.
The Z-scheme heterojunction is usually divided into direct Z-scheme and indirect Z-scheme heterojunctions, as shown in Figure 6e,f. The fundamental differences between direct Z-scheme and indirect Z-scheme heterojunctions primarily lie in the charge transfer pathways, semiconductor configurations, and the separation efficiency of photogenerated carriers. In the direct Z-scheme, two semiconductor materials achieve direct electron-hole pair recombination through intimate interfacial contact (without any additional mediators), forming an internal electric field-driven Z-scheme pathway. This mechanism offers high charge transfer efficiency (minimizing energy loss caused by mediators) and relies on the intrinsic band alignment of the semiconductors. Moreover, the direct Z-scheme features a simple structure and lower fabrication complexity. In contrast, the indirect Z-scheme requires electron mediators (e.g., noble metal nanoparticles or carbon materials) acting as “electron bridges” to facilitate indirect charge recombination. This process depends on the redox cycling of the mediators and often necessitates additional electrolyte components. While the mediators can enhance charge transfer efficiency, they may also introduce additional complexity to the system.
Direct Z-scheme heterojunctions are formed by direct contact between two semiconductors, where the two semiconductor materials are in close contact to form stronger interactions, which helps reduce the resistance to carrier transfer [107]. For example, BiOI quantum dots and g-C3N4 were fabricated at room temperature via the precursor in situ transformation method. By reducing the size of BiOI to the quantum scale, a complete intimate contact between BiOI and g-C3N4 was achieved, which in turn improved the catalytic activity [93]. Compared with the direct Z-type heterojunction of TiO2/g-C3N4, the indirect Z-scheme heterojunction of TiO2/BC/g-C3N4 uses BC as an electron mediator, which facilitates the separation of electron-hole pairs and reduces the loss in the electron transfer process. During the experiments, the efficient removal of Cr(VI) was realized by enhancing the interfacial mass transfer and accelerating the electron transfer in TiO2/BC/g-C3N4. The presence of a fully fixed Z-scheme connection structure can effectively solve the charge separation problem of the traditional II connection structure. Such a composite structure allows the use of narrow band gap semiconductor pairs without a loss in photoelectron and hole-strong redox capacity [108,109,110], preventing self-corrosion of certain unstable semiconductors by transferring the associated reduced electron/oxide holes. D.M. Chen et al. [111] prepared a AgBr/g-C3N4 Z photocatalyst for Ag/AgBr/g-C3N4 reduction, in which Ag nanoparticles acted as electron mediators to facilitate electron transfer from AgBr to g-C3N4. S.S. Lu et al. [112] prepared Bi/γ-Bi2O3/EtCN (O-doped g-C3N4), and an all-solid Z-type heterojunction was formed between γ-Bi2O3 and EtCN. The enhancement of the composite activity is mainly due to the modification of Bi after the formation of Z-type heterojunction, which broadens the light absorption range of the catalyst and promotes the effective separation of electron-hole pairs.
The composite materials with Z-scheme heterojunction, such as g-C3N4/Bi2WO3 [113], g-C3N4/Ag3PO4 [114], and g-C3N4/Ag3VO4 [115], usually have suitable electronic structures and good photochemical stability. The composite structure formed by mixing and calcining g-C3N4 with semiconductors has a good enhancement effect on photocatalytic performance [116,117]. The matched electronic structure of the two semiconductors and the combination of g-C3N4 with BiOX at low temperatures can improve the visible-light photocatalytic performance [118]. Under visible light, both types of semiconductors produce photogenerated charge carriers that can quickly separate photogenerated electrons from holes. Photogenerated charge carriers in a novel ternary heterojunction structure of g-C3N4/Bi2MoO6/Bi are transferred via an indirect Z-order structure. The synergistic effect of the Z-type heterojunction structure and the SPR effect of the doped Bi resulted in stronger light absorption and lower photogenerated charge carrier binding of the material (Figure 11) [119]. Z.L. Huang [120] fabricated Z-scheme two-dimensional/one-dimensional g-C3N4/MoO3-x composites, which exhibited strong visible and near-infrared absorption.

3.2.6. S-Scheme Heterojunction

In the field of photocatalysis, direct Z-scheme and S-scheme heterojunctions exhibit significant differences in their electron transfer and energy band mechanisms. The direct Z-scheme heterojunction forms a “Z”-shaped pathway relying on the band potential difference between semiconductors, requiring no mediator and featuring a simple structure, though it has high requirements for interfacial contact and is suitable for scenarios with clear band matching, such as CO2 reduction. In contrast, the S-scheme heterojunction is more distinctive; it drives the selective recombination of electrons and holes through a built-in electric field, preferentially consuming low-energy carriers to form an “S”-shaped transfer pathway. By relying on regulation of the Fermi levels of n-type and p-type semiconductors, it not only significantly improves charge separation efficiency and broadens the photoresponse range but also retains strong redox capabilities while efficiently adapting to complex catalytic needs such as water splitting and deep mineralization of pollutants. Despite higher requirements for preparation processes, the S-scheme heterojunction is becoming a new hotspot in the research and application of photocatalysis by virtue of its unique carrier regulation mechanism.
As shown in Figure 8, this heterojunction is mainly constructed from a reduced-type semiconductor photocatalyst (RP) with smaller work functions and higher Fermi energy levels and an oxidized-type semiconductor photocatalyst (OP) with larger work functions and lower Fermi energy levels through a staggered approach. When observing the structural schematic, it is similar to the Type Ⅱ heterojunction. However, in a typical Type Ⅱ heterojunction, photogenerated electrons and holes are accumulated in the conductive and valence bands of the RP and OP, respectively, leading to a weakened redox capacity. In contrast, in S-scheme heterojunctions, the effective electrons and holes are preserved, and the meaningless photogenerated carriers are recombined. This is named for the electron transfer schematic, which resembles a step. G. Huang et al. [121] synthesized stable ZnO/g-C3N4 S-scheme heterojunctions through simple solvent evaporation and high-temperature thermal polymerization reactions. Under the same experimental conditions, the photocatalytic degradation of pyridine via pure g-C3N4 and a ZnO/g-C3N4 heterojunction was 65.8% and 98.9%, respectively. A. Lsulmi et al. prepared an S-scheme AgI/g-C3N4 heterojunction with 96% photocatalytic degradation of RhB dye. The AgI nanoparticles degraded 40% of the RhB under the same conditions, and the pristine g-C3N4 was even less efficient [122].
Generally, in combination with a suitable semiconductor, a lower energy level semiconductor can pull the excited electrons of the graphite carbon nitride into its own conduction band, which can use this energy not only to excite the photo drive of the semiconductor but also reduce the electron-hole recombination rate of graphite carbon nitride and improve the absorption and utilization of visible light, thereby enhancing the photocatalytic ability. For Type I heterojunctions, electrons and holes in two semiconductors migrate in the same direction. In contrast, Type II heterojunctions allow electrons and holes to migrate to different semiconductors, resulting in space charge separation [123]. Combining p-type semiconductors with n-type semiconductors leads to the redistribution of electrons within the catalytic system, which reduces the compounding rate of photogenerated charges and enhances the photocatalytic performance of the catalysts. A Schottky junction is a close contact between the interface of a metal and a semiconductor, where the semiconductor bends upward for the sake of the Fermi energy levels between them, and the electrons in the conduction band move toward the metal. The Schottky junction can effectively separate the photogenerated charges, extend the lifetime of the photogenerated charges, and improve the photocatalytic performance of the photocatalytic material [124]. Conventional heterojunction has the advantages of expanding the range of photo-response and promoting carrier separation, but it has the problem of an insufficient oxidation-reduction capacity. The Z-scheme heterojunction has many advantages over it, including effective separation of electron-hole pairs, reduction of compounding chances, retention of strong oxidation-reduction active sites, broadening of the photo-response range, and an improvement in photocatalytic activity. In conventional Z-scheme heterojunctions, charge transfer between the two semiconductors is achieved using redox electron mediator pairs, and these ions can only achieve sufficient migration rates in a solution. This limits the system to the reaction of the solution phase. S-scheme heterojunctions achieve spatial separation of semiconductor photogenerated electrons and holes with a guaranteed strong redox capacity through three factors: a built-in electric field, energy band bending, and electrostatic interactions. Some applications of S-scheme heterojunctions have been described in this paper, including photocatalytic hydrogen production, carbon dioxide reduction, and pollutant degradation [125,126]. However, as a new heterojunction system, S-scheme heterojunctions have been studied less.

4. Application of Graphitic Carbon Nitride in Photocatalytic and Photoelectrocatalytic Degradation

Since the beginning of this century, photocatalytic research results have emerged and shown broad development [127,128]. Scientists have researched and developed various photocatalysts [129,130,131]. In the last 12 years, two-dimensional g-C3N4 has attracted much attention for applications in photocatalytic environmental remediation and solar energy conversion to fuel [132]. When irradiated with light, the g-C3N4 material mixes with the aqueous solution to form hydroxyl radicals, which have a strong oxidizing ability and decompose organic pollutants, realizing a reduction in the number of carbon atoms in the molecules of organic compounds, reducing the molecular weight, and even decomposing them into carbon dioxide and water.

4.1. Methodological Principles of Photocatalysis and Photoelectrocatalysis

Photocatalysis is an efficient oxidation technology that has emerged in recent years. In short, photocatalytic oxidation technology is an advanced redox technology that uses light energy. The reaction equation is simply described as follows:
Semiconductor (SC) + hv → h+ + e
Different types of semiconductors have different optical properties and colors, and these properties depend on the size of the band gap. The band gap is usually between 1.5 eV and 3.0 eV. Different band gap energies also reveal the difference in the absorption wavelengths, which directly leads to a difference in photocatalytic capacity [133,134].
PEC is an effective strategy for improving oxidation efficiency by utilizing photoelectric collaborative processes instead of traditional photocatalysis and electrocatalysis. As a new advanced oxidation technology, it combines the advantages of photocatalysis and electrocatalysis. By applying low auxiliary pressure to the electrode, photogenerated electrons are promoted, which are captured by the anode, transferred through the external circuit, and quickly reacted at the cathode, improving the key problem of easy recombination of e-h+ in photocatalysis [135]. Studies have confirmed that PEC technology exhibits far superior catalytic performance to single catalytic methods under identical conditions [136,137]. As shown in Figure 12, the PEC system consists of a photocatalytic anode and an electrode cathode. Electrode cathodes can be composed of various photocatalysts. Wires then connect the cathode and anode to form a closed-loop system. The core of the system is a photocatalytic semiconductor material. The photocatalytic anode generates electrons and holes when exposed to light. The electrons generated by light are transferred to the cathode through an external circuit. Under anaerobic conditions, they can react with hydrogen ions transferred from the anode to form H2. In the presence of oxygen, oxygen can be reduced to water. Therefore, the photo hole on the photo anode surface will oxidize and decompose the pollutants in the anode chamber, realizing the closed circuit and recovering the electric energy while eliminating the pollutants.

4.2. Relationship Between g-C3N4 Modification and Photo Response

The photoanode has always been a hot research topic as a key element of the photocatalytic and photoelectrocatalytic reaction system. Researchers are paying more attention to developing and improving photocatalysts for photoanodes [138]. g-C3N4 is attracting attention as a CN-based material. However, the photocatalytic stability of g-C3N4 is greatly limited due to its short life and the weak redox ability of the photo-generating carriers [139]. The g-C3N4 composite structure can transfer photogenerated electrons or holes to semiconductors, metals, oxides, and other materials [140,141], which enhances the photoelectrocatalytic reaction and efficiently prevents the generation of composite photogenerated carriers. g-C3N4-modified material catalysts have four advantages:
  • Being combined with a co-catalyst
An appropriate co-catalyst combined with g-C3N4 can increase the active site, reduce the overpotential, and improve the photocatalytic performance [142].
2.
Enhanced light absorption
g-C3N4 can act as a photocatalyst under visible light, and its modified materials can further enhance its light absorption. For example, assisted by nitrogen defects, g-C3N4 undergoes a pressure-driven structural transition (PIA) from a graphitic to an amorphous phase. Under high pressures, the narrowing of the band gap enhances light absorption in the visible range [143].
3.
Improvement in transmission efficiency
The separation of the photogenerated carrier can effectively separate the electron hole to improve transport efficiency.
4.
Good stability
Stable, non-toxic, and non-polluting g-C3N4 has good thermal and chemical stability. g-C3N4 and its modified materials are capable of stable performance at high temperatures. It is only at over 600 °C that the thermal stability begins to decrease.
Elemental doping and composite modification can significantly improve the visible-light reactivity of g-C3N4 by disrupting the intra-layer hydrogen bonding and enhancing the visible-light photocatalytic performance [144]. For example, the introduction of the S element replaces the edge N atom of sp2 hybridization, breaks the intra-layer hydrogen bond, and greatly improves the carbon atom density of g-C3N4, which not only promotes the n→π* electronic transition but also broadens the visible-light response range and lowers the potential barrier, which is conducive to photogenerated charge-carrier migration and separation and photogenerated electron trapping [145], with absorbance in the visible region for lithium-doped and silver-doped g-C3N4-modified composites, inhibition of photogenerated electron-hole pair recombination, and improved generation of reactive oxide species [146]. Constructing a heterojunction expands the range of light absorption and realizes different effects depending on the material, expanding the application of g-C3N4 materials. For example, by constructing heterojunctions using bismuth-based (BiOX, X = F, Cl, Br, and I) halogen oxides, it is possible to convert UV-active BiOF into visible-active BiOF, resulting in improved efficiency of the applied bias photon current and significant chemical stability over long periods [147].

4.3. Photocatalytic Performance of g-C3N4 and Its Modified Material Under Different Light Sources

g-C3N4 and its modified materials have shown great potential in the application of photocatalysis degradation under different light sources [148]. As we all know, the full spectrum contains the spectral curves of ultraviolet rays (UV), visible light (Vis), and near-infrared light (NIR). Compared with traditional TiO2 photocatalysts, g-C3N4 has a wider absorption spectrum. It has the natural advantage of being able to act as a photocatalyst under normal visible light without the need for UV light. However, the application of g-C3N4 under more light source conditions is still worth investigating.

4.3.1. UV Photocatalytic Degradation

UV is a general term for radiation in the electromagnetic spectrum with wavelengths from 0.01 to 0.40 μm (between the violet end of visible light and X-rays). UV light has a high energy level and can trigger reactions in a wide range of molecules or ions. Upon UV irradiation, atmospheric oxygen molecules decompose to generate free oxygen radicals. As free oxygen radicals carry unpaired electrons (in an unstable state), they readily combine with oxygen molecules to form ozone.Ozone has a strong oxidizing effect on exhaust gas molecules and is, therefore, highly effective in treating volatile organic gases (VOCs) [149]. Some researchers have also found that the UV 3D electrode’s electro-Fenton removal rate is consistently higher for dye wastewater than the 3D electrode’s electro-Fenton removal rate. UV-catalyzed H2O2 degradation of dyes is more effective under acidic conditions. The UV/H2O2 system promotes the oxidation of organic matter through hydrogen extraction, electron transfer, and radical addition [150,151]. Under UV irradiation, the photocatalytic degradation of RhB by TiO2/ZnO-g-C3N4 prepared by H.X. Jing [152] was superior to that of other light sources up to 99.6%. J.G. Kim et al. [153] prepared iron-modified g-C3N4 (FG materials) for the removal of arsenate (As(V)) and arsenite (As(III)) from water. The cerium fluoride (CeF3) semiconductor CeF3 has upconversion properties for converting Vis and NIR light into UV light. The CeF3/g-C3N4 heterojunction is capable of photocatalytic degradation of dibenzothiophene (DBT) [154]. The synergistic action of CeF3 and g-C3N4 achieves up to 84.2% desulfurization of DBT in 3 h. This composite has potential application in deep desulfurization.

4.3.2. Visible-Light Photocatalytic Degradation

Visible-light photocatalysis is a robust emerging area for organic transformation and is of great value [155]. Since the pioneering study of photocatalysis of g-C3N4 under visible-light irradiation by X.C. Wang et al. in 2009 [156], g-C3N4 has been recognized as a semiconductor with a narrow band gap (~2.7 eV) and visible light activity (~460 nm) [157,158]. The photocatalytic activity of Bi2MoO6/g-C3N4 composites under visible light significantly improved after introducing Bi2MoO6. After a certain time of visible-light irradiation, the degradation efficiency of organic pollutants was 93.88%, which was 3.9 times higher than that of g-C3N4 [159]. M. Sheydaei et al. [160] prepared visible-light active g-C3N4/Ce-ZnO/Ti nanocomposites and found that the visible-light photoelectrocatalytic process was able to degrade cefixime in aqueous solution with the highest efficiency. The WO3/g-C3N4 prepared by D.M. Li et al. had a porous and wrinkled structure. Its light absorption coefficient was larger than that of g-C3N4, and its band gap was 0.3 eV, which is smaller than that of g-C3N4. Therefore, the recombination rate of photogenerated electron-hole pairs of WO3/g-C3N4 was greatly reduced, and the visible light-induced catalytic activity was enhanced. The removal rate of RhB can reach 97.3% [161].

4.3.3. Near-Infrared Photocatalytic Degradation

Infrared light makes up more than 50% of the Sun’s spectrum. At present, the application research of photocatalysts is mainly focused on the ultraviolet and visible regions, while reports on the near-infrared region are relatively few in number. For efficient solar fuel conversion, researchers need to scrutinize the extension of light absorbance into the near-infrared (NIR) region. Theoretically, NIR-responsive photocatalysts for diffuse solar energy can be utilized to alleviate energy crises and environmental pollution [162].
A series of studies has shown that extending the photo response into the infrared optical region is possible by preparing heterojunctions with noble metals. The p-n heterojunction constructed using g-C3N4 and Ag2O not only extends the absorption spectrum into the near-infrared region but also inhibits the recombination of photogenerated carriers, thus further improving the catalytic degradation performance of Ag2O in the near infrared [163]. F. Wu et al. found that p doping can extend the response spectrum of g-C3N4 from 450 nm to 1000 nm, which produces a significant near-infrared response during PEC [164]. Furthermore, the Ag2O/Ag2S2O7 heterojunction catalyst with a full spectral response, prepared by H. Li et al. [165], can effectively improve the activity and stability of Ag2O in the UV, visible, and near-infrared regions. UCNPs can convert near-infrared photons into ultraviolet (UV) and visible light, enabling the utilization of near-infrared light. S.Y. Lee combined g-C3N4 with upconversion nanoparticles (UCNPs). The use of an NIR light source enhanced the photocatalytic activity of UCNP and g-C3N4 nanocomposites for efficient degradation of methyl orange (Figure 13) [166].

4.4. Application of g-C3N4 and Its Modified Materials in Photoelectrocatalysis Under Different Light Sources

Photoelectrocatalysis is an effective strategy for improving oxidation efficiency. In contrast to photocatalysis, PEC can promote electrochemical oxidation of pollutants at the anode and reduction reactions at the cathode by generating a bias voltage. In PEC technology, the fixation of g-C3N4 as a matrix material can minimize potential secondary contamination through material leaching and avoid additional post-processing steps. At present, in the photoelectrocatalytic system, the most commonly used light sources are xenon lamps (simulating sunlight, emitting light as a continuous spectrum from 200 nm to 2000 nm, and having energy distribution) and UV-visible light. Meanwhile, g-C3N4 can activate molecular oxygen and generate superoxide radicals for the photocatalytic conversion of organic functional groups and photocatalytic degradation of organic pollutants, which is more suitable for the treatment of indoor air pollution and the degradation of organic matter rather than TiO2 [167].
The PEC system prepared by impregnating g-C3N4 film onto ITO glass and irradiating it with a 500 W xenon lamp with a bias voltage of 2.5 V was able to increase the mineralization rate of phenol from 37% to 89% in a much shorter period of time. W. Zhen et al. [168] prepared a g-C3N4/TiO2 film photoanode, and the mineralization of phenol was only increased to 100% after 1.5 h. A g-C3N4/TiO2-NTs electrode was synthesized by Y. Zhang et al. [169], and the g-C3N4/TiO2-NTs electrodes showed higher TOC mineralization of phenol and benzyl alcohol than the TiO2-NTs electrodes within 2 h. The g-C3N4/TiO2-NTs electrodes can be used for the simultaneous removal of Cr(VI) from wastewater under visible-light irradiation. As shown in Figure 14, hexavalent chromium, phenol, or other organic pollutants in wastewater can be removed simultaneously under visible-light irradiation.
It is worth noting that the performance of the PEC system is highly dependent on the properties and type of the conductive substrate as well as the photocatalytic activity, especially under visible-light irradiation. It is an excellent method for constructing a composite photoanode by coupling g-C3N4 with a visible-light active semiconductor with a matching band structure, which can enhance light absorption and charge transfer through a highly interactive interface. The PEC system investigated using g-C3N4/Ag/AgCl/BiVO4 photoanode materials showed high effectiveness in treating wastewater of the pollutants diphenyl ketone and Escherichia coli [170]. Ag/AgCl media not only increase the light absorption and electron-hole transfer pathways of the g-C3N4/BiVO4 heterojunction but also promote high photoanode recovery through their polarization and plasma effects. In a study by C. Murugan [171], g-C3N4-BiVO4 direct Z-type heterojunctions were synthesized using a hydrothermal method, and Figure 15 shows a schematic diagram of their photoelectrocatalytic overall water separation. The results confirmed that the CSI response of BiVO4 photoanodes containing g-C3N4 was improved under AM 1.5 G (100 mWcm−2) light. The formation of g-C3N4-BiVO4 dielectric-free direct Z-type heterojunctions enhances interfacial charge transfer and inhibits charge recombination.
The g-C3N4/CNTs/Al2O3 photoanode was successfully integrated into the membrane filtration of the PEC system [172]. The conductive carbon nanotube layer and the visible-light responsive g-C3N4 layer were sequentially coated on the Al2O3 membrane carrier, and the composite membrane with a photo adsorption side length of 500 nm and high PEC stability was fabricated. In continuous flow mode, the phenol removal rate was up to 94% from 10 to about 60 min. L. Zhang et al. [173] used MnOx/g-C3N4 nanocomposites as a photoanode to decompose chloric acid (CA). Superoxide radicals and positive holes play a leading role in the PEC degradation of CA. The photoanode MnOx/g-C3N4-10 showed the highest photocatalytic performance under acidic conditions. Therefore, in a PEC system containing a g-C3N4-based anode, pollutant removal occurs at both electrodes. The photogenerated electrons separated from the g-C3N4-based photoanode, which are moved to the counter electrode through an external circuit, can reduce O2 to form reactive oxygen species and further oxidize pollutants. That aside, the holes in the photoanode also can directly oxidize pollutants.

5. Conclusions and Prospective

This article reviewed the modification of g-C3N4 and summarized its role in photocatalysis and photoelectrocatalysis and its catalytic degradation applications under different light sources. The modification and optimization of g-C3N4 can start from syntheszing a single material or a series of modifications to the material. The photodegradation performance of g-C3N4 photocatalysts can be improved by enlarging the effective reaction sites for photocatalysis, prolonging the lifecycle of photogenerated electron holes, and enhancing the visible-light responsiveness of the catalysts. In recent years, the modification of g-C3N4-based photo- or photoelectrocatalytic materials has broadened the range of light absorption and increased the absorption and utilization of light by the catalysts. Attempts have been made to apply them to the treatment of various organic pollutants, realizing the efficient degradation of organic pollutants in water. Despite the extensive academic research on and achievements with g-C3N4-based materials, there are still many challenges to be faced, such as the following:
  • Although the photoelectrocatalytic activity of g-C3N4-based photoelectrocatalytic materials has been significantly improved, it still cannot meet the production demand. For the final application of g-C3N4 photocatalysts in real pollution treatment, research needs to consider the simplicity and environmental safety of the synthesis method, as well as the recovery and recycling of catalysts.
  • There is a lack of systematic understanding of how modifying g-C3N4-based photocatalysts affects catalytic activity. Therefore, it is necessary to explore the modification mechanism of g-C3N4, elucidate the electron transfer pathways inside the catalysts, and improve the controllability of the modification of g-C3N4-based materials.
  • More kinds of pollutants with different properties were selected for photo- and photoelectrocatalytic experiments, and the effect of the solution injection volume on degradation efficiency should not be neglected so as to provide ideas and references for the development of more efficient g-C3N4-based photo- and photoelectrocatalysts.
Currently, the research on g-C3N4 is gradually deepening and has moved from the theoretical development stage to the practical application stage.

Funding

This work was supported by the Natural Science Foundation of Guangxi province (2024GXNSFDA010015), Guangxi Key Technologies R&D Program (No. GuikeAB24010213), Guilin Key Technologies R&D Program (No. Shike (2024)17), and the Research Foundation of Key Laboratory of New Processing Technology for Nonferrous Metal & Materials of the Ministry of Education and Guangxi Key Laboratory of Optical and Electronic Materials and Devices (20AA17, 22KF-23).

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Li, Z.; Xu, X.; Sheng, X.; Lin, P.; Tang, J.; Pan, L.; Kaneti, Y.V.; Yang, T.; Yamauchi, Y. Solar-Powered Sustainable Water Production: State-of-the-Art Technologies for Sunlight–Energy–Water Nexus. ACS Nano 2021, 15, 12535–12566. [Google Scholar] [CrossRef] [PubMed]
  2. Lin, J.; Ye, W.; Xie, M.; Seo, D.H.; Luo, J.; Wan, Y.; Van der Bruggen, B. Environmental impacts and remediation of dye-containing wastewater. Nat. Rev. Earth Environ. 2023, 4, 785–803. [Google Scholar] [CrossRef]
  3. Mojiri, A.; Zhou, J.L.; Robinson, B.; Ohashi, A.; Ozaki, N.; Kindaichi, T.; Farraji, H.; Vakili, M. Pesticides in aquatic environments and their removal by adsorption methods. Chemosphere 2020, 253, 126646. [Google Scholar] [CrossRef]
  4. Nguyen, M.K.; Lin, C.; Bui, X.T.; Rakib, M.R.J.; Nguyen, H.L.; Truong, Q.M.; Hoang, H.G.; Tran, H.-T.; Malafaia, G.; Idris, A.M. Occurrence and fate of pharmaceutical pollutants in wastewater: Insights on ecotoxicity, health risk, and state–of–the-art removal. Chemosphere 2024, 354, 141678. [Google Scholar] [CrossRef] [PubMed]
  5. Dong, Z.; Zhou, R.; Xiong, L.; Li, H.; Liu, Q.; Zheng, L.; Guo, Z.; Deng, Z. Preparation of a Ti0.7W0.3O2/TiO2 nanocomposite interfacial photocatalyst and its photocatalytic degradation of phenol pollutants in wastewater. Nanoscale Adv. 2020, 2, 425–437. [Google Scholar] [CrossRef]
  6. Zhou, Q.; Yang, N.; Li, Y.; Ren, B.; Ding, X.; Bian, H.; Yao, X. Total concentrations and sources of heavy metal pollution in global river and lake water bodies from 1972 to 2017. Glob. Ecol. Conserv. 2020, 22, e00925. [Google Scholar] [CrossRef]
  7. Dong, F.; Zhang, Y.X.; Zhang, S. Editorial: Photocatalysis for Environmental Applications. Front. Chem. 2019, 7. [Google Scholar] [CrossRef]
  8. Zhang, X.Y.; Yu, W.C.; Guo, Y.J.; Li, S.L.; Chen, Y.R.; Wang, H.; Bian, Z.Y. Recent advances in photoelectrocatalytic advanced oxidation processes: From mechanism understanding to catalyst design and actual applications. Chem. Eng. J. 2023, 455, 140801. [Google Scholar] [CrossRef]
  9. Li, D.; Qu, J. The progress of catalytic technologies in water purification: A review. J. Environ. Sci. 2009, 21, 713–719. [Google Scholar] [CrossRef]
  10. Galushchinskiy, A.; González-Gómez, R.; McCarthy, K.; Farràs, P.; Savateev, A. Progress in Development of Photocatalytic Processes for Synthesis of Fuels and Organic Compounds under Outdoor Solar Light. Energy Fuels 2022, 36, 4625–4639. [Google Scholar] [CrossRef]
  11. He, Y.; Yin, L.; Yuan, N.; Zhang, G. Adsorption and activation, active site and reaction pathway of photocatalytic CO2 reduction: A review. Chem. Eng. J. 2024, 481, 148754. [Google Scholar] [CrossRef]
  12. Mishra, S.; Acharya, L.; Marandi, B.; Sanjay, K.; Acharya, R. Boosted photocatalytic accomplishment of 3D/2D hierarchical structured Bi4O5I2/g-C3N4 p-n type direct Z-scheme heterojunction towards synchronous elimination of Cr(VI) and tetracycline. Diam. Relat. Mater. 2024, 142, 110834. [Google Scholar] [CrossRef]
  13. Acharya, R.; Mishra, S. Visible light sensitized graphitic carbon nitride based semiconducting materials for photoelectrochemical water splitting. Mater. Today Proc. 2023. [Google Scholar] [CrossRef]
  14. Zhu, P.J.; Yan, Y.N.; Zhou, Y.; Qi, Z.J.; Li, Y.F.; Chen, C.M. Thermal Properties of Graphene and Graphene-Based Nanocomposites: A Review. ACS Appl. Nano Mater. 2024, 7, 8377–9815. [Google Scholar] [CrossRef]
  15. Liang, F.; Zhu, Y. Enhancement of mineralization ability for phenol via synergetic effect of photoelectrocatalysis of g-C3N4 film. Appl. Catal. B Environ. 2016, 180, 324–329. [Google Scholar] [CrossRef]
  16. Paul, S.; Panja, S.; Hazra, N.; Gayen, K.; Banerjee, A. Carbon Dot as Visible-Light Photoredox Catalysts for a Myriad of Organic Transformations. J. Org. Chem. 2024, 89, 91–100. [Google Scholar] [CrossRef]
  17. Xie, S.Y.; Li, X.; Zheng, H.W.; Feng, L.; Khan, S. Research Progress of Organic Carbon Nanotubes-Modified Metal-Composite Photocatalytic Materials in Water Treatment. Mini-Rev. Org. Chem. 2022, 19, 898–905. [Google Scholar] [CrossRef]
  18. Xiangjuan, C.; Huan, W.; Weijia, A.; Li, L.; Wenquan, C. Study on Photoelectrocatalysis of Organic Carbon Materials. Prog. Chem. 2022, 34, 2361–2372. [Google Scholar] [CrossRef]
  19. Zhao, G.; Yang, H.C.; Liu, M.Q.; Xu, X.J. Metal-Free Graphitic Carbon Nitride Photocatalyst Goes Into Two-Dimensional Time. Front. Chem. 2018, 6, 551. [Google Scholar] [CrossRef]
  20. 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]
  21. Ong, W.J.; Tan, L.L.; Ng, Y.H.; Yong, S.T.; Chai, S.P. Graphitic Carbon Nitride (g-C3N4)-Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer To Achieving Sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  22. Wudil, Y.S.; Ahmad, U.F.; Gondal, M.A.; Al-Osta, M.A.; Almohammedi, A.; Sa’id, R.S.; Hrahsheh, F.; Haruna, K.; Mohamed, M.J.S. Tuning of graphitic carbon nitride (g-C3N4) for photocatalysis: A critical review. Arab. J. Chem. 2023, 16, 104542. [Google Scholar] [CrossRef]
  23. Zou, X.; Sun, Z.; Hu, Y.H. g-C3N4-based photoelectrodes for photoelectrochemical water splitting: A review. J. Mater. Chem. A 2020, 8, 21474–21502. [Google Scholar] [CrossRef]
  24. 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]
  25. Schwarzer, A.; Saplinova, T.; Kroke, E. Tri-s-triazines (s-heptazines)-From a “mystery molecule” to industrially relevant carbon nitride materials. Coord. Chem. Rev. 2013, 257, 2032–2062. [Google Scholar] [CrossRef]
  26. Lotsch, B.V.; Schnick, W. New Light on an Old Story: Formation of Melam during Thermal Condensation of Melamine. Chem. Eur. J. 2007, 13, 4956–4968. [Google Scholar] [CrossRef] [PubMed]
  27. Zhang, Y.; Fugane, K.; Mori, T.; Niu, L.; Ye, J. Wet chemical synthesis of nitrogen-doped graphene towards oxygen reduction electrocatalysts without high-temperature pyrolysis. J. Mater. Chem. 2012, 22, 6575–6580. [Google Scholar] [CrossRef]
  28. Zhang, Y.; Mori, T.; Ye, J. Polymeric Carbon Nitrides: Semiconducting Properties and Emerging Applications in Photocatalysis and Photoelectrochemical Energy Conversion. Sci. Adv. Mater. 2012, 4, 282–291. [Google Scholar] [CrossRef]
  29. Thomas, A.; Fischer, A.; Goettmann, F.; Antonietti, M.; Müller, J.-O.; Schlögl, R.; Carlsson, J. ChemInform Abstract: Graphitic Carbon Nitride Materials: Variation of Structure and Morphology and Their Use as Metal-Free Catalysts. J. Mater. Chem. 2008, 40, 4893–4908. [Google Scholar] [CrossRef]
  30. Wang, J.B.; Lei, J.L.; Wang, R.H. Diffraction-pattern calculation and phase identification of hypothetical crystalline C3N4. Phys. Rev. B 1998, 58, 11890–11895. [Google Scholar] [CrossRef]
  31. Yao, H.; Ching, W.Y. Optical properties of beta -C3N4 and its pressure dependence. Phys. Rev. B Condens. Matter 1994, 50, 11231–11234. [Google Scholar] [CrossRef] [PubMed]
  32. Luo, J.; Wen, B.; Melnik, R. Relative stability of nanosized β-C3N4 and graphitic C3N4 from first principles calculations. Phys. E Low-Dimens. Syst. Nanostructures 2012, 45, 190–193. [Google Scholar] [CrossRef]
  33. Liu, A.Y.; Cohen, M.L. Structural properties and electronic structure of low-compressibility materials: Beta -Si3N4 and hypothetical beta-C3N4. Phys. Rev. B Condens. Matter 1990, 41, 10727–10734. [Google Scholar] [CrossRef] [PubMed]
  34. Mattesini, M.; Matar, S.F.; Etourneau, J. Stability and electronic property investigations of the graphitic C3N4 system showing an orthorhombic unit cell. J. Mater. Chem. 2000, 10, 709–713. [Google Scholar] [CrossRef]
  35. Sahoo, S.; Acharya, R. An overview on recent developments in synthesis and molecular level structure of visible-light responsive g-C3N4 photocatalyst towards environmental remediation. Mater. Today Proc. 2021, 35, 150–155. [Google Scholar] [CrossRef]
  36. Acharya, R.; Parida, K. A review on TiO2/g-C3N4 visible-light- responsive photocatalysts for sustainable energy generation and environmental remediation. J. Environ. Chem. Eng. 2020, 8, 103896. [Google Scholar] [CrossRef]
  37. Takanabe, K.; Kamata, K.; Wang, X.C.; Antonietti, M.; Kubota, J.; Domen, K. Photocatalytic hydrogen evolution on dye-sensitized mesoporous carbon nitride photocatalyst with magnesium phthalocyanine. Phys. Chem. Chem. Phys. 2010, 12, 13020–13025. [Google Scholar] [CrossRef]
  38. Sohail, M.; Anwar, U.; Taha, T.A.; Qazi, H.I.A.; Al-Sehemi, A.G.; Ullah, S.; Algarni, H.; Ahmed, I.M.; Amin, M.A.; Palamanit, A.; et al. Nanostructured materials based on g-C3N4 for enhanced photocatalytic activity and potentials application: A review. Arab. J. Chem. 2022, 15, 104070. [Google Scholar] [CrossRef]
  39. Wang, Q.; Cao, L.; Liang, S.J.; Wu, W.; Wang, G.; Lee, C.H.; Ong, W.L.; Yang, H.Y.; Ang, L.K.; Yang, S.A.; et al. Efficient Ohmic contacts and built-in atomic sublayer protection in MoSi2N4 and WSi2N4 monolayers. Npj 2D Mater. Appl. 2021, 5, 71. [Google Scholar] [CrossRef]
  40. Zhang, Y.; Liu, S.; Ji, N.; Wei, L.; Liang, Q.; Li, J.; Tian, Z.; Su, J.; Chen, Q. Modulation of the electronic structure of metallic bismuth catalysts by cerium doping to facilitate electrocatalytic CO2 reduction to formate. J. Mater. Chem. A 2024, 12, 7528–7535. [Google Scholar] [CrossRef]
  41. Ma, D.; Zhang, Z.; Zou, Y.; Chen, J.; Shi, J.-W. The progress of g-C3N4 in photocatalytic H2 evolution: From fabrication to modification. Coord. Chem. Rev. 2024, 500, 215489. [Google Scholar] [CrossRef]
  42. Xu, J.; Zhang, L.; Shi, R.; Zhu, Y. Chemical exfoliation of graphitic carbon nitride for efficient heterogeneous photocatalysis. J. Mater. Chem. A 2013, 1, 14766–14772. [Google Scholar] [CrossRef]
  43. Tyborski, T.; Merschjann, C.; Orthmann, S.; Yang, F.; Lux-Steiner, M.C.; Schedel-Niedrig, T. Crystal structure of polymeric carbon nitride and the determination of its process-temperature-induced modifications. J. Phys. Condens. Matter 2013, 25, 395402. [Google Scholar] [CrossRef] [PubMed]
  44. Buriak, J.M.; Toro, C.; Choi, K.-S. Chemistry of Materials for Water Splitting Reactions. Chem. Mater. 2018, 30, 7325–7327. [Google Scholar] [CrossRef]
  45. Lotsch, B.V.; Döblinger, M.; Sehnert, J.; Seyfarth, L.; Senker, J.; Oeckler, O.; Schnick, W. Unmasking Melon by a Complementary Approach Employing Electron Diffraction, Solid-State NMR Spectroscopy, and Theoretical Calculations—Structural Characterization of a Carbon Nitride Polymer. Chem. A Eur. J. 2007, 13, 4969–4980. [Google Scholar] [CrossRef] [PubMed]
  46. Ouyang, K.; Xu, B.; Yang, C.; Wang, H.; Zhan, P.; Xie, S. Synthesis of a novel Z-scheme Ag/WO3/g-C3N4 nanophotocatalyst for degradation of oxytetracycline hydrochloride under visible light. Mater. Sci. Semicond. Process. 2022, 137, 106168. [Google Scholar] [CrossRef]
  47. Pan, T.; Chen, D.; Fang, J.; Wu, K.; Feng, W.; Zhu, X.; Fang, Z. Facile synthesis of iron and cerium co-doped g-C3N4 with synergistic effect to enhance visible-light photocatalytic performance. Mater. Res. Bull. 2020, 125, 110812. [Google Scholar] [CrossRef]
  48. Yang, Y.; Niu, W.; Dang, L.; Mao, Y.; Wu, J.; Xu, K. Recent Progress in Doped g-C3N4 Photocatalyst for Solar Water Splitting: A Review. Front. Chem. 2022, 10, 955065. [Google Scholar] [CrossRef]
  49. Hu, X.; Guo, R.T.; Lin, Z.D.; Bi, Z.X.; Chen, X.; Wang, J.; Pan, W. Construction of Carbon Dot-Modified g-C3N4/BiOIO3 Z-Scheme Heterojunction for Boosting Photocatalytic CO2 Reduction under Full Spectrum Light. ACS Sustain. Chem. Eng. 2022, 10, 11143–11153. [Google Scholar] [CrossRef]
  50. Jiang, H.Y.; Li, Y.; Wang, D.H.; Hong, X.D.; 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]
  51. Cui, W.; Chen, P.; Chen, L.C.; Li, J.Y.; Zhou, Y.; Dong, F. Alkali/alkaline-earth metal intercalated g-C3N4 induced charge redistribution and optimized photocatalysis: Status and challenges. J. Phys. Energy 2021, 3, 032008. [Google Scholar] [CrossRef]
  52. Bai, L.Q.; Huang, H.W.; Yu, S.X.; Zhang, D.Y.; Huang, H.T.; Zhang, Y.H. Role of transition metal oxides in g-C3N4-based heterojunctions for photocatalysis and supercapacitors. J. Energy Chem. 2022, 64, 214–235. [Google Scholar] [CrossRef]
  53. Phoon, B.L.; Ong, C.C.; Lee, K.C.; Pan, G.T.; Leo, B.F.; Chong, S.; Pan, K.L. Co-doped, tri-doped, and rare-earth-doped g-C3N4 for photocatalytic applications: State-of-the-Art. Catalysts 2022, 12, 586. [Google Scholar] [CrossRef]
  54. Arunachalapandi, M.; Roopan, S.M. Environment friendly g-C3N4-Based catalysts and their recent strategy in organic transformations. High. Energy Chem. 2022, 56, 73–90. [Google Scholar] [CrossRef]
  55. Zhang, H.; Tang, Y.Q.; Liu, Z.X.; Zhu, Z.; Tang, X.; Wang, Y.M. Study on optical properties of alkali metal doped g-C3N4 and their photocatalytic activity for reduction of CO2. Chem. Phys. Lett. 2020, 751, 137467. [Google Scholar] [CrossRef]
  56. Ye, S.; Qiu, L.G.; Yuan, Y.P.; Zhu, Y.J.; Xia, J.; Zhu, J.F. Facile fabrication of magnetically separable graphitic carbon nitride photocatalysts with enhanced photocatalytic activity under visible light. J. Mater. Chem. A 2013, 1, 3008–3015. [Google Scholar] [CrossRef]
  57. Mugaka, B.P.; Zhang, S.; Li, R.Q.; Ma, Y.; Wang, B.; Hong, J.; Hu, Y.H.; Ding, Y.; Xia, X.H. One-Pot Preparation of Peptide-Doped Metal-Amino Acid Framework for General Encapsulation and Targeted Delivery. ACS Appl. Mater. Interfaces 2021, 13, 11195–11204. [Google Scholar] [CrossRef] [PubMed]
  58. Deng, P.; Xiong, J.; Lei, S.; Wang, W.; Ou, X.; Xu, Y.; Xiao, Y.; Cheng, B. Nickel formate induced high-level in situ Ni-doping of g-C3N4 for a tunable band structure and enhanced photocatalytic performance. J. Mater. Chem. A 2019, 7, 22385–22397. [Google Scholar] [CrossRef]
  59. Ma, J.; Jia, N.; Shen, C.; Liu, W.; Wen, Y. Stable cuprous active sites in Cu+-graphitic carbon nitride: Structure analysis and performance in Fenton-like reactions. J. Hazard. Mater. 2019, 378, 120782. [Google Scholar] [CrossRef]
  60. Ruan, L.; Xu, G.; Gu, L.; Li, C.; Zhu, Y.; Lu, Y. The physical properties of Li-doped g-C3N4 monolayer sheet investigated by the first-principles. Mater. Res. Bull. 2015, 66, 156–162. [Google Scholar] [CrossRef]
  61. Ta, Q.T.H.; Namgung, G.; Noh, J.S. Facile synthesis of porous metal-doped ZnO/g-C3N4 composites for highly efficient photocatalysts. J. Photochem. Photobiol. A Chem. 2019, 368, 110–119. [Google Scholar] [CrossRef]
  62. Li, W.; Zhang, H.; Chen, P.; Yao, J.; Dong, X. One-Step Hydrothermal Deposition of Ag-Doped g-C3N4-TiO2 Nanocomposites on Cotton Fabric Surface with Enhanced Photocatalytic Activity. Fibers Polym. 2023, 24, 575–588. [Google Scholar] [CrossRef]
  63. Ji, S.; Yang, Y.; Zhou, Z.; Li, X.; Liu, Y. Photocatalysis-Fenton of Fe-doped g-C3N4 catalyst and its excellent degradation performance towards RhB. J. Water Process Eng. 2021, 40, 101804. [Google Scholar] [CrossRef]
  64. Tang, C.S.; Cheng, M.; Lai, C.; Li, L.; Yang, X.F.; Du, L.; Zhang, G.X.; Wang, G.F.; Yang, L. Recent progress in the applications of non-metal modified graphitic carbon nitride in photocatalysis. Coord. Chem. Rev. 2023, 474, 214846. [Google Scholar] [CrossRef]
  65. Starukh, H.; Praus, P. Doping of Graphitic Carbon Nitride with Non-Metal Elements and Its Applications in Photocatalysis. Catalysts 2020, 10, 1119. [Google Scholar] [CrossRef]
  66. Kuchmiy, S.Y.; Stroyuk, O.L. Photocatalytic Fixation of Molecular Nitrogen in Systems Based on Graphite-Like Carbon Nitride: A Review. Theor. Exp. Chem. 2021, 57, 85–112. [Google Scholar] [CrossRef]
  67. Inagaki, M.; Tsumura, T.; Kinumoto, T.; Toyoda, M. Graphitic carbon nitrides (g-C3N4) with comparative discussion to carbon materials. Carbon 2019, 141, 580–607. [Google Scholar] [CrossRef]
  68. Wang, Y.; Wang, H.; Chen, F.; Cao, F.; Zhao, X.; Meng, S.; Cui, Y. Facile synthesis of oxygen doped carbon nitride hollow microsphere for photocatalysis. Appl. Catal. B Environ. 2017, 206, 417–425. [Google Scholar] [CrossRef]
  69. Wang, Y.; Shen, S. Progress and Prospects of Non-Metal Doped Graphitic Carbon Nitride for Improved Photocatalytic Performances. Acta Phys. Chim. Sin. 2020, 36, 1905080. [Google Scholar] [CrossRef]
  70. Arumugam, M.; Tahir, M.; Praserthdam, P. Effect of nonmetals (B, O, P, and S) doped with porous g-C3N4 for improved electron transfer towards photocatalytic CO2 reduction with water into CH4. Chemosphere 2022, 286, 131765. [Google Scholar] [CrossRef]
  71. Liu, G.; Niu, P.; Sun, C.; Smith, S.C.; Chen, Z.; Lu, G.Q.; Cheng, H.M. Unique electronic structure induced high photoreactivity of sulfur-doped graphitic C3N4. J. Am. Chem. Soc. 2010, 132, 11642–11648. [Google Scholar] [CrossRef] [PubMed]
  72. Dangwang Dikdim, J.M.; Gong, Y.; Noumi, G.B.; Sieliechi, J.M.; Zhao, X.; Ma, N.; Yang, M.; Tchatchueng, J.B. Peroxymonosulfate improved photocatalytic degradation of atrazine by activated carbon/graphitic carbon nitride composite under visible light irradiation. Chemosphere 2019, 217, 833–842. [Google Scholar] [CrossRef]
  73. Zhang, J.; Zhao, X.; Wang, Y.; Gong, Y.; Cao, D.; Qiao, M. Peroxymonosulfate-enhanced visible light photocatalytic degradation of bisphenol A by perylene imide-modified g-C3N4. Appl. Catal. B Environ. 2018, 237, 976–985. [Google Scholar] [CrossRef]
  74. Yang, B.; Zhou, H.; Zhang, X.; Zhao, M. Electron spin-polarization and band gap engineering in carbon-modified graphitic carbon nitrides. J. Mater. Chem. C 2015, 3, 10886–10891. [Google Scholar] [CrossRef]
  75. Ling, F.; Li, W.; Ye, L. The synergistic effect of non-metal doping or defect engineering and interface coupling on the photocatalytic property of g-C3N4: First-principle investigations. Appl. Surf. Sci. 2019, 473, 386–392. [Google Scholar] [CrossRef]
  76. 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]
  77. Li, Z.; Raziq, F.; Liu, C.; Bai, L.; Jing, L. Surface-engineering strategies for g-C3N4 as efficient visible-light photocatalyst. Curr. Opin. Green Sustain. Chem. 2017, 6, 57–62. [Google Scholar] [CrossRef]
  78. Que, M.D.; Cai, W.H.; Chen, J.; Zhu, L.L.; Yang, Y.W. Recent advances in g-C3N4 composites within four types of heterojunctions for photocatalytic CO2 reduction. Nanoscale 2021, 13, 6692–6712. [Google Scholar] [CrossRef] [PubMed]
  79. Zhao, C.; Gong, X.; Wang, S.; Jiang, W.; Xuan, S. Shear Stiffening Gels for Intelligent Anti-impact Applications. Cell Rep. Phys. Sci. 2020, 1, 100266. [Google Scholar] [CrossRef]
  80. Medina-Llamas, M.; Speltini, A.; Profumo, A.; Panzarea, F.; Milella, A.; Fracassi, F.; Listorti, A.; Malavasi, L. Preparation of Heterojunctions Based on Cs3Bi2Br9 Nanocrystals and g-C3N4 Nanosheets for Photocatalytic Hydrogen Evolution. Nanomaterials 2023, 13, 263. [Google Scholar] [CrossRef]
  81. Yang, Y.; Zhu, J.F.; He, Y.F.; Li, M.; Liu, Y.; Chen, M.M.; Cao, D.W. Charge transfer in photocatalysis of direct Z-scheme g-C3N4-based ferroelectric heterojunction. J. Alloys Compd. 2022, 893, 162270. [Google Scholar] [CrossRef]
  82. Ren, Y.; Zeng, D.; Ong, W.J. Interfacial engineering of graphitic carbon nitride (g-C3N4)-based metal sulfide heterojunction photocatalysts for energy conversion: A review. Chin. J. Catal. 2019, 40, 289–319. [Google Scholar] [CrossRef]
  83. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-Scheme heterojunction photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  84. Acharya, R.; Pati, S.; Parida, K. A review on visible light driven spinel ferrite-g-C3N4 photocatalytic systems with enhanced solar light utilization. J. Mol. Liq. 2022, 357, 119105. [Google Scholar] [CrossRef]
  85. Ji, X.Y.; Guo, R.T.; Tang, J.Y.; Lin, Z.D.; Yuan, Y.; Hong, L.F.; Pan, W.G. Fabrication of a ternary NiS/ZnIn2S4/g-C3N4 photocatalyst with dual charge transfer channels towards efficient H2 evolution. J. Colloid Interface Sci. 2022, 618, 300–310. [Google Scholar] [CrossRef] [PubMed]
  86. Li, Y.; Wang, X.; Huo, H.; Li, Z.; Shi, J. A novel binary visible-light-driven photocatalyst type-I CdIn2S4/g-C3N4 heterojunctions coupling with H2O2: Synthesis, characterization, photocatalytic activity for Reactive Blue 19 degradation and mechanism analysis. Colloids Surf. A Physicochem. Eng. Asp. 2020, 587, 124322. [Google Scholar] [CrossRef]
  87. Zhang, X.; Yang, P.; Jiang, S.P. The edge-epitaxial growth of yellow g-C3N4 on red g-C3N4 nanosheets with superior photocatalytic activities. Chem. Commun. 2021, 57, 3119–3122. [Google Scholar] [CrossRef] [PubMed]
  88. Xiong, W.; Dong, Y.; Pan, A. Fabricating a type II heterojunction by growing lead-free perovskite Cs2AgBiBr6in situ on graphite-like g-C3N4 nanosheets for enhanced photocatalytic CO2 reduction. Nanoscale 2023, 15, 15619–15625. [Google Scholar] [CrossRef] [PubMed]
  89. Cao, H.; Yan, Y.; Wang, Y.; Chen, F.F.; Yu, Y. Dual role of g-C3N4 microtubes in enhancing photocatalytic CO2 reduction of Co3O4 nanoparticles. Carbon 2023, 201, 415–424. [Google Scholar] [CrossRef]
  90. Hu, S.; Zhai, W.; Chen, F.; He, Q. Improvement of separation, transfer and redox ability arising from P–N heterojuncted composite phosphate-group-intercalated g-C3N4 /Bi2WO6. Mater. Today Phys. 2024, 40, 101311. [Google Scholar] [CrossRef]
  91. Chen, Y.; Ren, X.; Wang, X.; Tian, Z.; Yang, X.; Lu, J.; Bai, H.; Jiao, T.; Huang, H.; Hu, J. Construction of Ag decorated P-doped g-C3N4 nanosheets Schottky junction via silver mirror reaction for enhanced photocatalytic activities. Int. J. Hydrog. Energy 2022, 47, 250–263. [Google Scholar] [CrossRef]
  92. Li, J.; Zhao, L.; Wang, S.; Li, J.; Wang, G.; Wang, J. In situ fabrication of 2D/3D g-C3N4/Ti3C2 (MXene) heterojunction for efficient visible-light photocatalytic hydrogen evolution. Appl. Surf. Sci. 2020, 515, 145922. [Google Scholar] [CrossRef]
  93. He, R.; Cheng, K.; Wei, Z.; Zhang, S.; Xu, D. Room-temperature in situ fabrication and enhanced photocatalytic activity of direct Z-scheme BiOI/g-C3N4 photocatalyst. Appl. Surf. Sci. 2019, 465, 964–972. [Google Scholar] [CrossRef]
  94. Chen, M.; Wang, G.; Dai, J.; Li, H.; Deng, N. Indirect Z-scheme TiO2/BC/g-C3N4 for efficient photocatalytic reduction of Cr(VI) in aqueous solution. J. Chem. Technol. Biotechnol. 2024, 99, 415–425. [Google Scholar] [CrossRef]
  95. Bai, K.; Cui, Z.; Li, E.; Ding, Y.; Zheng, J.; Liu, C.; Zheng, Y. Electronic and optical characteristics of GaS/g-C3N4 van der Waals heterostructures: Effects of biaxial strain and vertical electric field. Vacuum 2020, 180, 109562. [Google Scholar] [CrossRef]
  96. Yang, C.; Yang, J.; Liu, S.; Zhao, M.; Duan, X.; Wu, H.; Liu, L.; Liu, W.; Li, J.; Ren, S.; et al. Constructing C–O bridged CeO2/g-C3N4 S-scheme heterojunction for methyl orange photodegradation: Experimental and theoretical calculation. J. Environ. Manag. 2023, 335, 117608. [Google Scholar] [CrossRef]
  97. Chen, T.; Zhang, J.; Ge, H.; Li, M.; Li, Y.; Liu, B.; Duan, T.; He, R.; Zhu, W. Efficient extraction of uranium in organics-containing wastewater over g-C3N4/GO hybrid nanosheets with type-II band structure. J. Hazard. Mater. 2020, 384, 121383. [Google Scholar] [CrossRef]
  98. Dong, F.; Zhao, Z.; Xiong, T.; Ni, Z.; Zhang, W.; Sun, Y.; Ho, W.-K. In Situ Construction of g-C3N4/g-C3N4 Metal-Free Heterojunction for Enhanced Visible-Light Photocatalysis. ACS Appl. Mater. Interfaces 2013, 5, 11392–11401. [Google Scholar] [CrossRef]
  99. 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]
  100. Liu, J.; Zhang, X.; Zhong, Q.; Li, J.; Wu, H.; Zhang, B.; Jin, L.; Tao, H.B.; Liu, B. Electrostatic self-assembly of a AgI/Bi2Ga4O9 p–n junction photocatalyst for boosting superoxide radical generation. J. Mater. Chem. A 2020, 8, 4083–4090. [Google Scholar] [CrossRef]
  101. Lei, J.; Gu, X.; Xiao, P.; Ding, G.; Yang, Y.; Fu, X.; Long, B.; Chen, S.; Meng, S. Fabrication of 2D/2D BiOBr/g-C3N4 with efficient photocatalytic activity and clarification of its mechanism. Phys. Chem. Chem. Phys. 2022, 24, 19806–19816. [Google Scholar] [CrossRef] [PubMed]
  102. Zhong, H.; Yang, C.; Fan, L.; Fu, Z.; Yang, X.; Wang, X.; Wang, R. Dyadic promotion of photocatalytic aerobic oxidation via the Mott–Schottky effect enabled by nitrogen-doped carbon from imidazolium-based ionic polymers. Energy Environ. Sci. 2019, 12, 418–426. [Google Scholar] [CrossRef]
  103. Yang, L.; Charnas, A.; Qiu, G.; Lin, Y.-M.; Lu, C.-C.; Tsai, W.; Paduano, Q.; Snure, M.; Ye, P.D. How Important Is the Metal–Semiconductor Contact for Schottky Barrier Transistors: A Case Study on Few-Layer Black Phosphorus? ACS Omega 2017, 2, 4173–4179. [Google Scholar] [CrossRef]
  104. Yan, F.; Wang, Y.; Zhang, J.; Lin, Z.; Zheng, J.; Huang, F. Schottky or Ohmic Metal–Semiconductor Contact: Influence on Photocatalytic Efficiency of Ag/ZnO and Pt/ZnO Model Systems. ChemSusChem 2014, 7, 101–104. [Google Scholar] [CrossRef]
  105. Hu, T.; Yang, Y.; Dai, K.; Zhang, J.; Liang, C. A novel Z-scheme Bi2MoO6/BiOBr photocatalyst for enhanced photocatalytic activity under visible light irradiation. Appl. Surf. Sci. 2018, 456, 473–481. [Google Scholar] [CrossRef]
  106. Li, H.; Hu, T.; Zhang, R.; Liu, J.; Hou, W. Preparation of solid-state Z-scheme Bi2MoO6/MO (MCu, Co3/4, or Ni) heterojunctions with internal electric field-improved performance in photocatalysis. Appl. Catal. B Environ. 2016, 188, 313–323. [Google Scholar] [CrossRef]
  107. Mishra, S.; Acharya, R. Recent updates in modification strategies for escalated performance of Graphene/MFe2O4 heterostructured photocatalysts towards energy and environmental applications. J. Alloys Compd. 2023, 960, 170576. [Google Scholar] [CrossRef]
  108. Ghosh, U.; Pal, A. Graphitic carbon nitride based Z scheme photocatalysts: Design considerations, synthesis, characterization and applications. J. Ind. Eng. Chem. 2019, 79, 383–408. [Google Scholar] [CrossRef]
  109. Liao, G.F.; Li, C.X.; Li, X.Z.; Fang, B.Z. Emerging polymeric carbon nitride Z-scheme systems for photocatalysis. Cell Rep. Phys. Sci. 2021, 2, 100355. [Google Scholar] [CrossRef]
  110. Jourshabani, M.; Lee, B.K.; Shariatinia, Z. From Traditional Strategies to Z-scheme Configuration in Graphitic Carbon Nitride Photocatalysts: Recent Progress and Future Challenges. Appl. Catal. B-Environ. 2020, 276, 119157. [Google Scholar] [CrossRef]
  111. Chen, D.; Wang, Z.; Du, Y.; Yang, G.; Ren, T.; Ding, H. In situ ionic-liquid-assisted synthesis of plasmonic photocatalyst Ag/AgBr/g-C3N4 with enhanced visible-light photocatalytic activity. Catal. Today 2015, 258, 41–48. [Google Scholar] [CrossRef]
  112. Lu, S.; Wu, T.; Liu, Y.; Luo, H.; Jiang, F.; Nie, X.; Chen, H. All-solid Z-scheme Bi/γ-Bi2O3/O-doped g-C3N4 heterojunction with Bi as electron shuttle for visible-light photocatalysis. J. Alloys Compd. 2022, 911, 164980. [Google Scholar] [CrossRef]
  113. Zhang, Y.; Chai, C.; Zhang, X.; Liu, J.; Duan, D.; Fan, C.; Wang, Y. Construction of Pt-decorated g-C3N4/Bi2WO6 Z-scheme composite with superior solar photocatalytic activity toward rhodamine B degradation. Inorg. Chem. Commun. 2019, 100, 81–91. [Google Scholar] [CrossRef]
  114. Cui, Y.; Yang, L.; Meng, M.; Zhang, Q.; Li, B.; Wu, Y.; Zhang, Y.; Lang, J.; Li, C. Facile preparation of antifouling g-C3N4/Ag3PO4 nanocomposite photocatalytic polyvinylidene fluoride membranes for effective removal of rhodamine B. Korean J. Chem. Eng. 2019, 36, 236–247. [Google Scholar] [CrossRef]
  115. Zhu, T.; Song, Y.; Ji, H.; Xu, Y.; Song, Y.; Xia, J.; Yin, S.; Li, Y.; Xu, H.; Zhang, Q.; et al. Synthesis of g-C3N4/Ag3VO4 composites with enhanced photocatalytic activity under visible light irradiation. Chem. Eng. J. 2015, 271, 96–105. [Google Scholar] [CrossRef]
  116. Darkwah, W.K.; Ao, Y.H. Mini Review on the Structure and Properties (Photocatalysis), and Preparation Techniques of Graphitic Carbon Nitride Nano-Based Particle, and Its Applications. Nanoscale Res. Lett. 2018, 13, 388. [Google Scholar] [CrossRef]
  117. Zhang, X.B.; Song, H.Y.; Sun, C.Y.; Chen, C.X.; Han, F.Q.; Li, X.F. Photocatalytic oxidative desulfurization and denitrogenation of fuels over sodium doped graphitic carbon nitride nanosheets under visible light irradiation. Mater. Chem. Phys. 2019, 226, 34–43. [Google Scholar] [CrossRef]
  118. Sridharan, K.; Shenoy, S.; Kumar, S.G.; Terashima, C.; Fujishima, A.; Pitchaimuthu, S. Advanced Two-Dimensional Heterojunction Photocatalysts of Stoichiometric and Non-Stoichiometric Bismuth Oxyhalides with Graphitic Carbon Nitride for Sustainable Energy and Environmental Applications. Catalysts 2021, 11, 426. [Google Scholar] [CrossRef]
  119. Li, N.; Gao, H.; Wang, X.; Zhao, S.; Lv, D.; Yang, G.; Gao, X.; Fan, H.; Gao, Y.; Ge, L. Novel indirect Z-scheme g-C3N4/Bi2MoO6/Bi hollow microsphere heterojunctions with SPR-promoted visible absorption and highly enhanced photocatalytic performance. Chin. J. Catal. 2020, 41, 426–434. [Google Scholar] [CrossRef]
  120. Huang, Z.L.; Liu, J.C.; Zong, S.; Wang, X.Y.; Chen, K.X.; Liu, L.L.; Fang, Y.X. Fabrication of graphitic carbon Nitride/Nonstoichiometric molybdenum oxide nanorod composite with the nonmetal plasma enhanced photocatalytic hydrogen evolution activity. J. Colloid Interface Sci. 2022, 606, 848–859. [Google Scholar] [CrossRef]
  121. Huang, G.; Liu, S.; Tian, C.; Tao, Y. Construction of S scheme ZnO/g-C3N4 heterojunction for the removal of pyridine from coal chemical wastewater. Opt. Mater. 2024, 150, 115288. [Google Scholar] [CrossRef]
  122. Alsulmi, A.; Hussein, I.A.; Nasherty, M.; Hesham, M.; Soltan, A.; Messih, M.F.A.; Ahmed, M.A. Sonochemical Fabrication of S-Scheme AgI/g-C3N4 Heterojunction for Efficient Photocatalytic Degradation of RhB Dye. J. Inorg. Organomet. Polym. Mater. 2024, 34, 640–654. [Google Scholar] [CrossRef]
  123. Mohammad, A.; Ahmad, K.; Qureshi, A.; Tauqeer, M.; Mobin, S.M. Zinc oxide-graphitic carbon nitride nanohybrid as an efficient electrochemical sensor and photocatalyst. Sens. Actuators B Chem. 2018, 277, 467–476. [Google Scholar] [CrossRef]
  124. Huang, H.; Liu, C.; Ou, H.; Ma, T.; Zhang, Y. Self-sacrifice transformation for fabrication of type-I and type-II heterojunctions in hierarchical BixOyIz/g-C3N4 for efficient visible-light photocatalysis. Appl. Surf. Sci. 2019, 470, 1101–1110. [Google Scholar] [CrossRef]
  125. Raza, A.; Haidry, A.A.; Amin, T.; Hussain, A.A.; Shah, S.A.M.H.; Ahsan, M. Boosting the water splitting and hydrogen production of S-scheme fabricated porous g-C3N4 modified with CuO. Diam. Relat. Mater. 2024, 141, 110703. [Google Scholar] [CrossRef]
  126. Kong, X.; Fan, J.; Feng, B.; Li, J.; Yang, G.; Xue, C. Carbon dots-triggered the fabrication of miniature g-C3N4/CDs/WO3 S-scheme heterojunction for efficient CO2 photoreduction. Chem. Eng. J. 2023, 476, 146774. [Google Scholar] [CrossRef]
  127. Wan, Y.; Du, S.; Lu, C.; Ren, K.; Shi, B.; Liu, S.; Li, C.; Dou, W.; Fang, P.; Ye, N. Metallic CuS decorated CdS nanowires for efficient photocatalytic H2 evolution under visible-light irradiation. J. Alloys Compd. 2021, 871, 159461. [Google Scholar] [CrossRef]
  128. Garrido, I.; Flores, P.; Hellín, P.; Vela, N.; Navarro, S.; Fenoll, J. Solar reclamation of agro-wastewater polluted with eight pesticides by heterogeneous photocatalysis using a modular facility. A Case Study. Chemosphere 2020, 249, 126156. [Google Scholar] [CrossRef]
  129. Li, J.; Dong, X.; Sun, Y.; Jiang, G.; Chu, Y.; Lee, S.C.; Dong, F. Tailoring the rate-determining step in photocatalysis via localized excess electrons for efficient and safe air cleaning. Appl. Catal. B Environ. 2018, 239, 187–195. [Google Scholar] [CrossRef]
  130. Zeng, Q.; Wang, X.; Xie, X.; Mahmood, A.; Lu, G.; Wang, Y.; Sun, J. Band bending of TiO2 induced by O-xylene and acetaldehyde adsorption and its effect on the generation of active radicals. J. Colloid Interface Sci. 2020, 572, 374–383. [Google Scholar] [CrossRef]
  131. Li, K.; Lu, X.; Zhang, Y.; Liu, K.; Huang, Y.; Liu, H. Bi3TaO7/Ti3C2 heterojunctions for enhanced photocatalytic removal of water-borne contaminants. Environ. Res. 2020, 185, 109409. [Google Scholar] [CrossRef]
  132. Prabhu, S.; Cindrella, L.; Kwon, O.J.; Mohanraju, K. Photoelectrochemical and photocatalytic activity of TiO2-WO3 heterostructures boosted by mutual interaction. Mater. Sci. Semicond. Process. 2018, 88, 10–19. [Google Scholar] [CrossRef]
  133. Yan, S.C.; Li, Z.S.; Zou, Z.G. Photodegradation performance of g-C3N4 fabricated by directly heating melamine. Langmuir 2009, 25, 10397–10401. [Google Scholar] [CrossRef]
  134. Mamba, G.; Mishra, A.K. Graphitic carbon nitride (g-C3N4) nanocomposites: A new and exciting generation of visible light driven photocatalysts for environmental pollution remediation. Appl. Catal. B-Environ. 2016, 198, 347–377. [Google Scholar] [CrossRef]
  135. Arai, S. Chemistry of chromian spinel in volcanic rocks as a potential guide to magma chemistry. Mineral. Mag. 1992, 56, 173–184. [Google Scholar] [CrossRef]
  136. Zhang, M.; Yang, Y.; An, X.; Hou, L.A.A. critical review of g-C3N4-based photocatalytic membrane for water purification. Chem. Eng. J. 2021, 412, 128663. [Google Scholar] [CrossRef]
  137. Tang, L.; Liu, L.; Yang, F. FeMoO4-graphene oxide photo-electro-catalyst for berberine removal and hydrogen evolution. Int. J. Hydrog. Energy 2019, 44, 19755–19761. [Google Scholar] [CrossRef]
  138. Gao, B.; Peng, C.; Chen, G.Z.; Li Puma, G. Photo-electro-catalysis enhancement on carbon nanotubes/titanium dioxide (CNTs/TiO2) composite prepared by a novel surfactant wrapping sol–gel method. Appl. Catal. B Environ. 2008, 85, 17–23. [Google Scholar] [CrossRef]
  139. Ye, S.; Chen, Y.; Yao, X.; Zhang, J. Simultaneous removal of organic pollutants and heavy metals in wastewater by photoelectrocatalysis: A review. Chemosphere 2021, 273, 128503. [Google Scholar] [CrossRef]
  140. Wang, W.; Zhou, C.; Yang, Y.; Zeng, G.; Zhang, C.; Zhou, Y.; Yang, J.; Huang, D.; Wang, H.; Xiong, W.; et al. Carbon nitride based photocatalysts for solar photocatalytic disinfection, can we go further? Chem. Eng. J. 2021, 404, 126540. [Google Scholar] [CrossRef]
  141. Xu, Y.; Song, J.; Chen, F.; Wang, X.F.; Yu, H.G.; Yu, J.G. Amorphous Ti(IV)-modified Bi2WO6 with enhanced photocatalytic performance. RSC Adv. 2016, 6, 65902–65910. [Google Scholar] [CrossRef]
  142. Zhao, Y.S.; Zhang, L.J.; Qi, J.; Jin, Q.; Lin, K.F.; Wang, D. Graphdiyne with Enhanced Ability for Electron Transfer. Acta Phys.-Chim. Sin. 2018, 34, 1048–1060. [Google Scholar] [CrossRef]
  143. An, P.F.; Zhu, W.H.; Qiao, L.Y.; Sun, S.C.; Xu, Y.Y.; Jiang, D.L.; Chen, M.; Meng, S.C. 0D ultrafine ruthenium quantum dot decorated 3D porous graphitic carbon nitride with efficient charge separation and appropriate hydrogen adsorption capacity for superior photocatalytic hydrogen evolution. Dalton Trans. 2021, 50, 2414–2425. [Google Scholar] [CrossRef] [PubMed]
  144. Cheng, P.; Yao, D.; Yan, J.; Ye, T.; Liu, H.; Zeng, H.; Pan, X.; Zhang, G.; Ding, J. Pressure-Optimized Band Gap and Enhanced Photoelectric Response of Graphitic Carbon Nitride with Nitrogen Vacancies. Phys. Rev. Appl. 2023, 19, 024048. [Google Scholar] [CrossRef]
  145. Li, Y.; Zhu, S.; Liang, Y.; Li, Z.; Wu, S.; Chang, C.; Luo, S.; Cui, Z. One-step synthesis of Mo and S co-doped porous g-C3N4 nanosheets for efficient visible-light photocatalytic hydrogen evolution. Appl. Surf. Sci. 2021, 536, 147743. [Google Scholar] [CrossRef]
  146. Yang, C.; Zhang, S.; Huang, Y.; Lv, K.; Fang, S.; Wu, X.; Li, Q.; Fan, J. Sharply increasing the visible photoreactivity of g-C3N4 by breaking the intralayered hydrogen bonds. Appl. Surf. Sci. 2020, 505, 144654. [Google Scholar] [CrossRef]
  147. Ma, W.; Wang, X.; Zhang, F.; Fei, X.; Zhang, X.; Ma, H.; Dong, X. Synergetic effect of Li doping and Ag deposition for enhanced visible light photocatalytic performance of g-C3N4. Mater. Res. Bull. 2017, 86, 72–79. [Google Scholar] [CrossRef]
  148. Selvaraj, V.; Pandikumar, A. Turning UV Light-Active BiOF into Visible Light-Active BiOF by Forming a Heterojunction with g-C3N4 and Its Photoelectrochemical Water Splitting Performance in Reverse Osmosis-Rejected Wastewater. J. Phys. Chem. C 2022, 126, 79–90. [Google Scholar] [CrossRef]
  149. Gahlot, S.; Dappozze, F.; Mishra, S.; Guillard, C. High surface area g-C3N4 and g-C3N4-TiO2 photocatalytic activity under UV and Visible light: Impact of individual component. J. Environ. Chem. Eng. 2021, 9, 105587. [Google Scholar] [CrossRef]
  150. Wu, X.; Tan, Z.; Liu, R.; Liao, Z.; Ou, H. Gaseous products generated from polyethylene and polyethylene terephthalate during ultraviolet irradiation: Mechanism, pathway and toxicological analyses. Sci. Total Environ. 2023, 876, 162717. [Google Scholar] [CrossRef]
  151. Bezerra, K.C.H.; Fiaschitello, T.R.; Labuto, G.; Freeman, H.S.; Fragoso, W.D.; da Costa, S.M.; da Costa, S.A. Reuse of water from real reactive monochromic and trichromic wastewater for new cotton dyes after efficient treatment using H2O2 catalyzed by UV light. J. Environ. Chem. Eng. 2021, 9, 105731. [Google Scholar] [CrossRef]
  152. Asaithambi, P.; Govindarajan, R.; Yesuf, M.B.; Alemayehu, E. Removal of color, COD and determination of power consumption from landfill leachate wastewater using an electrochemical advanced oxidation processes. Sep. Purif. Technol. 2020, 233, 115935. [Google Scholar] [CrossRef]
  153. Hongxia, J.; Yanlin, G.; Longxiang, L.; Xu, W.; Wangjun, P. A new double Z-scheme TiO2/ZnO-g-C3N4 nanocomposite with enhanced photodegradation efficiency for Rhodamine B under sunlight. Environ. Prog. Sustain. Energy 2023, 42, e13968. [Google Scholar] [CrossRef]
  154. Lu, X.; Chen, F.; Qian, J.; Fu, M.; Jiang, Q.; Zhang, Q. Facile fabrication of CeF3/g-C3N4 heterojunction photocatalysts with upconversion properties for enhanced photocatalytic desulfurization performance. J. Rare Earths 2021, 39, 1204–1210. [Google Scholar] [CrossRef]
  155. Kim, J.G.; Kim, H.B.; Choi, J.H.; Baek, K. Bifunctional iron-modified graphitic carbon nitride (g-C3N4) for simultaneous oxidation and adsorption of arsenic. Environ. Res. 2020, 188, 109832. [Google Scholar] [CrossRef] [PubMed]
  156. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; 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] [CrossRef] [PubMed]
  157. Lang, X.J.; Chen, X.D.; Zhao, J.C. Heterogeneous visible light photocatalysis for selective organic transformations. Chem. Soc. Rev. 2014, 43, 473–486. [Google Scholar] [CrossRef]
  158. Liao, G.F.; Gong, Y.; Zhang, L.; Gao, H.Y.; Yang, G.J.; Fang, B.Z. Semiconductor polymeric graphitic carbon nitride photocatalysts: The "holy grail" for the photocatalytic hydrogen evolution reaction under visible light. Energy Environ. Sci. 2019, 12, 2080–2147. [Google Scholar] [CrossRef]
  159. Li, K.; Chen, M.; Chen, L.; Zhao, S.; Xue, W.; Han, Y. Investigating the Effect of Bi2MoO6/g-C3N4 Ratio on Photocatalytic Degradation of Sulfadiazine under Visible Light. Processes 2023, 11, 1059. [Google Scholar] [CrossRef]
  160. Sheydaei, M.; Ayoubi-Feiz, B.; Abbaszade-Fakhri, G. A visible-light active g-C3N4/Ce–ZnO/Ti nanocomposite for efficient photoelectrocatalytic pharmaceutical degradation: Modelling with artificial neural network. Process Saf. Environ. Prot. 2021, 149, 776–785. [Google Scholar] [CrossRef]
  161. Li, D.; Zhang, W.; Huang, Y.; Feng, H.; Wang, Z.; Yang, Z.; Chen, J.; Zhang, X.; Zhang, G.; Chen, Y. Visible light-induced catalytic performance of composite photocatalyst synthesized with nanomaterials WO3 and two-dimensional ultrathin g-C3N4. Water Sci. Technol. 2023, 88, 1910–1925. [Google Scholar] [CrossRef]
  162. Wang, J.; Wang, H.; Zuo, S.; Jin, X.; Zheng, B.; Deng, R.; Liu, W.; Wang, J. Synergistic effects of lanthanide surface adhesion and photon-upconversion for enhanced near-infrared responsive photodegradation of organic contaminants in wastewater. Environ. Sci. Nano 2020, 7, 3333–3342. [Google Scholar] [CrossRef]
  163. Liang, S.; Zhang, D.; Pu, X.; Yao, X.; Han, R.; Yin, J.; Ren, X. A novel Ag2O/g-C3N4 p-n heterojunction photocatalysts with enhanced visible and near-infrared light activity. Sep. Purif. Technol. 2019, 210, 786–797. [Google Scholar] [CrossRef]
  164. Wu, F.; Ma, Y.; Hu, Y.H. Near Infrared Light-Driven Photoelectrocatalytic Water Splitting over P-Doped g-C3N4. ACS Appl. Energy Mater. 2020, 3, 11223–11230. [Google Scholar] [CrossRef]
  165. Li, H.; Chen, T.; Wang, Y.; Tang, J.; Wang, Y.; Sang, Y.; Liu, H. Surface-sulfurized Ag2O nanoparticles with stable full-solar-spectrum photocatalytic activity. Chin. J. Catal. 2017, 38, 1063–1071. [Google Scholar] [CrossRef]
  166. Lee, S.Y.; Lee, G.; Jun, Y.S.; Park, Y.I. Visible/near-infrared driven highly efficient photocatalyst based on upconversion nanoparticles/g-C3N4 nanocomposite. Appl. Surf. Sci. 2020, 508, 144839. [Google Scholar] [CrossRef]
  167. Tian, H.; Liu, X.; Liang, Z.; Qiu, P.; Qian, X.; Cui, H.; Tian, J. Gold nanorods/g-C3N4 heterostructures for plasmon-enhanced photocatalytic H2 evolution in visible and near-infrared light. J. Colloid Interface Sci. 2019, 557, 700–708. [Google Scholar] [CrossRef] [PubMed]
  168. Wei, Z.; Liang, F.; Liu, Y.; Luo, W.; Wang, J.; Yao, W.; Zhu, Y. Photoelectrocatalytic degradation of phenol-containing wastewater by TiO2/g-C3N4 hybrid heterostructure thin film. Appl. Catal. B-Environ. 2017, 201, 600–606. [Google Scholar] [CrossRef]
  169. Zhang, Y.; Wang, Q.; Lu, J.; Wang, Q.; Cong, Y. Synergistic photoelectrochemical reduction of Cr(VI) and oxidation of organic pollutants by g-C3N4/TiO2-NTs electrodes. Chemosphere 2016, 162, 55–63. [Google Scholar] [CrossRef]
  170. Rather, R.A.; Lo, I.M.C. Photoelectrochemical sewage treatment by a multifunctional g-C3N4/Ag/AgCl/BiVO4 photoanode for the simultaneous degradation of emerging pollutants and hydrogen production, and the disinfection of E. coli. Water Res. 2020, 168, 115166. [Google Scholar] [CrossRef]
  171. Murugan, C.; Nataraj, R.A.; Kumar, M.P.; Ravichandran, S.; Pandikumar, A. Enhanced charge transfer process of bismuth vanadate interleaved graphitic carbon nitride nanohybrids in mediator-free direct Z scheme photoelectrocatalytic water splitting. Chemistryselect 2019, 4, 4653–4663. [Google Scholar] [CrossRef]
  172. Wang, X.; Wang, G.; Chen, S.; Fan, X.; Quan, X.; Yu, H. Integration of membrane filtration and photoelectrocatalysis on g-C3N4/CNTs/Al2O3 membrane with visible-light response for enhanced water treatment. J. Membr. Sci. 2017, 541, 153–161. [Google Scholar] [CrossRef]
  173. Zhang, L.; Zhang, M.; Song, X.; Wang, H.; Bian, Z. One-step synthesis of MnOx/g-C3N4 nanocomposites for enhancing the visible light photoelectrochemical oxidation performance. Chem. Eng. J. 2020, 399, 125825. [Google Scholar] [CrossRef]
Figure 1. Application fields of g-C3N4.
Figure 1. Application fields of g-C3N4.
Inorganics 13 00225 g001
Figure 2. Schematic diagram of g-C3N4 modification strategy.
Figure 2. Schematic diagram of g-C3N4 modification strategy.
Inorganics 13 00225 g002
Figure 3. Optimized structures of intrinsic g-C3N4 and alkali metal-doped g-C3N4: (a) intrinsic g-C3N4, (b) Li-doped g-C3N4, (c) Na-doped g-C3N4, (d) K-doped g-C3N44, (e) Rb-doped C3N4 (redrawn based on [55]).
Figure 3. Optimized structures of intrinsic g-C3N4 and alkali metal-doped g-C3N4: (a) intrinsic g-C3N4, (b) Li-doped g-C3N4, (c) Na-doped g-C3N4, (d) K-doped g-C3N44, (e) Rb-doped C3N4 (redrawn based on [55]).
Inorganics 13 00225 g003
Figure 4. Mechanism diagram for the degradation of rhodamine B by Cu+-graphite (copyright ScienceDirect, 2019) [59].
Figure 4. Mechanism diagram for the degradation of rhodamine B by Cu+-graphite (copyright ScienceDirect, 2019) [59].
Inorganics 13 00225 g004
Figure 5. Prediction schematic of band gap change in nonmetallic doped g-C3N4 (a) (redrawn based on [77]) and schematic illustration of the band structures of different types of g-C3N4 samples (b) [76]. VB-XPS = valence band X-ray photoelectron spectroscopy; MS = electrochemical analysis via Mott–Schottky plots (copyright ScienceDirect, 2017).
Figure 5. Prediction schematic of band gap change in nonmetallic doped g-C3N4 (a) (redrawn based on [77]) and schematic illustration of the band structures of different types of g-C3N4 samples (b) [76]. VB-XPS = valence band X-ray photoelectron spectroscopy; MS = electrochemical analysis via Mott–Schottky plots (copyright ScienceDirect, 2017).
Inorganics 13 00225 g005
Figure 6. Band structures of various types of heterojunctions in a photocatalytic hybrid nanocomposite: (a) Type I heterojunction, (b) Type II heterojunction, (c) p-n junction, (d) Schottky junction, (e) direct Z-scheme heterojunction (without an electron mediator), and (f) indirect Z-scheme (with an electron mediator). A, D, and Ef represent the electron acceptor, electron donor, and Fermi level, respectively (redrawn based on [82]).
Figure 6. Band structures of various types of heterojunctions in a photocatalytic hybrid nanocomposite: (a) Type I heterojunction, (b) Type II heterojunction, (c) p-n junction, (d) Schottky junction, (e) direct Z-scheme heterojunction (without an electron mediator), and (f) indirect Z-scheme (with an electron mediator). A, D, and Ef represent the electron acceptor, electron donor, and Fermi level, respectively (redrawn based on [82]).
Inorganics 13 00225 g006
Figure 7. S-scheme heterojunction mechanism diagram (redrawn based on [83]).
Figure 7. S-scheme heterojunction mechanism diagram (redrawn based on [83]).
Inorganics 13 00225 g007
Figure 8. Schematic illustration of photocatalytic H2 evolution over NiS/ZIS/UCN composite under UV–vis light irradiation (copyright ScienceDirect, 2022) [95].
Figure 8. Schematic illustration of photocatalytic H2 evolution over NiS/ZIS/UCN composite under UV–vis light irradiation (copyright ScienceDirect, 2022) [95].
Inorganics 13 00225 g008
Figure 9. The proposed charge transfer path over the tubular g-C3N4(TCN)/Co3O4 NP (copyright ScienceDirect, 2023) [89].
Figure 9. The proposed charge transfer path over the tubular g-C3N4(TCN)/Co3O4 NP (copyright ScienceDirect, 2023) [89].
Inorganics 13 00225 g009
Figure 10. Schematic diagram of the proposed mechanism for photocatalytic H2 production over g-C3N4/Ti3C2 composite (redrawn based on [92]).
Figure 10. Schematic diagram of the proposed mechanism for photocatalytic H2 production over g-C3N4/Ti3C2 composite (redrawn based on [92]).
Inorganics 13 00225 g010
Figure 11. Schematic photocatalytic mechanism of 0.4CN/BMO/9Bi composite during RhB degradation (redrawn based on [119]).
Figure 11. Schematic photocatalytic mechanism of 0.4CN/BMO/9Bi composite during RhB degradation (redrawn based on [119]).
Inorganics 13 00225 g011
Figure 12. Photoelectric catalytic PEC reaction mechanism (TiO2) (copyright ScienceDirect, 2019) [128].
Figure 12. Photoelectric catalytic PEC reaction mechanism (TiO2) (copyright ScienceDirect, 2019) [128].
Inorganics 13 00225 g012
Figure 13. Photocatalytic degradation of RhB with UCNPs/g-C3N4 (redrawn based on [166]).
Figure 13. Photocatalytic degradation of RhB with UCNPs/g-C3N4 (redrawn based on [166]).
Inorganics 13 00225 g013
Figure 14. The mechanism of charge separation and electron transfer in a g-C3N4/TiO2-NTs electrode under UV–visible-light irradiation (redrawn based on [169]).
Figure 14. The mechanism of charge separation and electron transfer in a g-C3N4/TiO2-NTs electrode under UV–visible-light irradiation (redrawn based on [169]).
Inorganics 13 00225 g014
Figure 15. Schematic of overall water splitting in a direct Z-type heterojunction of the g-C3N4-BiVO4 nanohybrid materials (redrawn based on [171]).
Figure 15. Schematic of overall water splitting in a direct Z-type heterojunction of the g-C3N4-BiVO4 nanohybrid materials (redrawn based on [171]).
Inorganics 13 00225 g015
Table 1. Comparison of modification methods.
Table 1. Comparison of modification methods.
StrategyPrincipleAdvantagesDisadvantages
DopingMetal dopingThe doped metal ions are positively charged and have strong interactions with the negatively charged C and N, which can form coordination bonds. This changes the lattice structure of g-C3N4.Alteration of the g-C3N4 lattice structure to reduce the band gap and expand the absorption range of visible light.Limited resources and high prices.
Non-metal dopingThe non-metallic elements themselves have high electro-negativity and ionization energies and can form covalent bonds with other compounds during the reaction process.Non-toxic and harmless, abundant sources, simple preparation process, good thermal and chemical stability, excellent light absorption performance, and adjustable band gap structure.Performance is not yet up to the performance requirements of noble metal-based catalysts.
HeterojunctionThe interface between two regions of different semiconductors with unequal band structures create interfacial band alignments.The heterojunction structure of g-C3N4 was constructed to be able to effectively inhibit carrier composite.The method for determining the type of heterojunction is complex and needs to be analyzed on an experimental basis.
Table 2. Examples of the performance of metal-doped g-C3N4.
Table 2. Examples of the performance of metal-doped g-C3N4.
Doped MetalPollutantsDegradation EfficiencyReferences
Alkali metal-doped g-C3N4Pure g-C3N4CO23.6 μmol g−1[55]
Li5.6 μmol g−1
Na7.4 μmol g−1
K9.8 μmol g−1
Rb12.1 μmol g−1
Ni-doping of
g-C3N4
Pure g-C3N4MO47.8 (140 min)[58]
Ni97.3 (90 min)
Cu-graphitic carbon nitrideCuRhB99.2 (60 min)[59]
ZnO/g-C3N4ZnOMB91 (20 min)[61]
Fe-g-C3N4Pure g-C3N4RhB69 (45 min)[63]
5% Fe92.9 (45 min)
10% Fe95.5 (45 min)
15% Fe90.5 (45 min)
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

Zhang, Y.; Lian, P.; Hao, X.; Zhang, L.; Yang, L.; Jiang, L.; Zhang, K.; Liao, L.; Qin, A. Modification Strategies of g-C3N4-Based Materials for Enhanced Photoelectrocatalytic Degradation of Pollutants: A Review. Inorganics 2025, 13, 225. https://doi.org/10.3390/inorganics13070225

AMA Style

Zhang Y, Lian P, Hao X, Zhang L, Yang L, Jiang L, Zhang K, Liao L, Qin A. Modification Strategies of g-C3N4-Based Materials for Enhanced Photoelectrocatalytic Degradation of Pollutants: A Review. Inorganics. 2025; 13(7):225. https://doi.org/10.3390/inorganics13070225

Chicago/Turabian Style

Zhang, Yijie, Peng Lian, Xinyu Hao, Li Zhang, Lihua Yang, Li Jiang, Kaiyou Zhang, Lei Liao, and Aimiao Qin. 2025. "Modification Strategies of g-C3N4-Based Materials for Enhanced Photoelectrocatalytic Degradation of Pollutants: A Review" Inorganics 13, no. 7: 225. https://doi.org/10.3390/inorganics13070225

APA Style

Zhang, Y., Lian, P., Hao, X., Zhang, L., Yang, L., Jiang, L., Zhang, K., Liao, L., & Qin, A. (2025). Modification Strategies of g-C3N4-Based Materials for Enhanced Photoelectrocatalytic Degradation of Pollutants: A Review. Inorganics, 13(7), 225. https://doi.org/10.3390/inorganics13070225

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop