Next Article in Journal
Physical and Electrical Properties of Silicon Nitride Thin Films with Different Nitrogen–Oxygen Ratios
Previous Article in Journal
Au-Co Alloy Nanoparticles Supported on ZrO2 as an Efficient Photocatalyst for the Deoxygenation of Styrene Oxide
Previous Article in Special Issue
Insights into CoFe2O4/Peracetic Acid Catalytic Oxidation Process for Iopamidol Degradation: Performance, Mechanisms, and I-DBP Formation Control
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Pre-, In-Process, and Post-Synthetic Strategies to Break the Surface Area Barrier in g-C3N4 for Energy Conversion and Environmental Remediation

1
College of Biological and Chemical Engineering, Qilu Institute of Technology, Jinan 250200, China
2
Key Laboratory of Superlight Materials & Surface Technology of Ministry of Education, Harbin Engineering University, Harbin 150001, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(13), 956; https://doi.org/10.3390/nano15130956
Submission received: 13 May 2025 / Revised: 15 June 2025 / Accepted: 17 June 2025 / Published: 20 June 2025

Abstract

Nanomaterials with large specific surface area (SSA) have emerged as pivotal platforms for energy storage and environmental remediation, primarily due to their enhanced active site exposure, improved mass transport capabilities, and superior interfacial reactivity. Among them, polymeric carbon nitride (g-C3N4) has garnered significant attention in energy and environmental applications owing to its visible-light-responsive bandgap (~2.7 eV), exceptional thermal/chemical stability, and earth-abundant composition. However, the practical performance of g-C3N4 is fundamentally constrained by intrinsic limitations, including its inherently low SSA (<20 m2/g via conventional thermal polymerization), rapid recombination of photogenerated carriers, and inefficient charge transfer kinetics. Notably, the theoretical SSA of g-C3N4 reaches 2500 m2/g, yet achieving this value remains challenging due to strong interlayer van der Waals interactions and structural collapse during synthesis. Recent advances demonstrate that state-of-the-art strategies can elevate its SSA to 50–200 m2/g. To break this surface area barrier, advanced strategies achieve SSA enhancement through three primary pathways: pre-treatment (molecular and supramolecular precursor design), in process (templating and controlled polycondensation), and post-processing (chemical exfoliation and defect engineering). This review systematically examines controllable synthesis methodologies for high-SSA g-C3N4, analyzing how SSA amplification intrinsically modulates band structures, extends carrier lifetimes, and boosts catalytic efficiencies. Future research should prioritize synergistic multi-stage engineering to approach the theoretical SSA limit (2500 m2/g) while preserving robust optoelectronic properties.

Graphical Abstract

1. Introduction

With the rapid development of global industrialization, energy shortages and environmental pollution have emerged as critical challenges worldwide. Addressing these issues requires dual approaches, including exploring alternative clean energy sources and developing advanced technologies for environmental remediation. In this context, semiconductor materials with tailored photoelectrochemical properties have attracted significant attention, owing to their ability to harness solar energy for sustainable applications [1,2,3,4]. Notably, nanostructured semiconductors with large specific surface areas (SSA) have been extensively investigated, as their enhanced surface reactivity, abundant active sites, and improved mass transport properties can substantially boost catalytic performance. Representative materials like activated carbon (with SSA up to 3000 m2/g) and graphene (featuring a 2D layered structure) exemplify how a high SSA enables exceptional adsorption capacity and electron transfer efficiency [5,6,7,8,9]. However, unlike conductive materials, semiconductors require precisely engineered band structures to balance light absorption and redox potential. Integrating the unique advantages of semiconductors with high SSA therefore represents a promising strategy to optimize energy use and tackle environmental crises.
Among semiconductor candidates, polymeric carbon nitride (g-C3N4), composed of earth-abundant carbon and nitrogen elements, has stood out as a metal-free photocatalyst since its discovery for water splitting in 2009 [10]. Its appropriate bandgap (~2.7 eV), visible-light responsiveness, and exceptional thermal/chemical stability (derived from triazine or 3-s-triazine ring units) have enabled widespread applications in photocatalytic hydrogen evolution, CO2 reduction, organic pollutant degradation, and so on [11,12,13,14,15,16]. Nevertheless, the practical performance of bulk g-C3N4 remains far below theoretical expectations due to inherent limitations [17] including low SSA (<20 m2/g) caused by strong interlayer van der Waals interactions and incomplete exfoliation during conventional thermal polymerization, rapid recombination of photogenerated carriers due to insufficient active sites and inefficient charge separation, and l.imited depth of light absorption resulting from particle aggregation. Intriguingly, the theoretical SSA of g-C3N4 nanosheet (up to 2500 m2/g) suggests immense potential for improvement through nanostructure engineering and surface modification, thereby addressing these bottlenecks [18].
Significant efforts have been devoted to enhancing the SSA of g-C3N4 via morphological control (e.g., porous architectures, nanotubes, nanosheets) and defect engineering [19,20,21,22]. These strategies not only increase the exposure of active sites but also modulate electronic structures to improve light absorption and charge transfer kinetics. For instance, 2D g-C3N4 nanosheets exhibit enhanced photocatalytic activity due to their ultrathin structure and shortened carrier migration paths [23,24]. However, critical trade-offs persist. Quantum size effects in nanostructured g-C3N4 may inadvertently widen the bandgap, compromising visible-light utilization, and excessive defects (e.g., pores, vacancies) can act as recombination centers, counteracting SSA benefits. Interplay between morphology, band structure, and defect density requires delicate balancing to optimize overall performance.
While prior reviews have comprehensively cataloged general modification strategies for g-C3N4, they have seldomed examined surface area engineering through the critical lens of process chronology. Specifically, the distinct roles of pre-treatment (precursor design), in-process (polymerization control), and post-processing (exfoliation/functionalization) strategies remain underexplored in correlating SSA enhancement with application-specific performance. This review bridges that gap by establishing a unified framework, systematically evaluating how interventions at each phase uniquely govern pore architecture and active site density to break the surface area barrier, enabling optimal energy use and environmental applications.
This review rigorously examines advances in high-SSA g-C3N4 synthesis through three chronologically defined pathways, focusing on pre-treatment strategies (precursor molecular engineering and supramolecular assembly), in-process modulation (templating and salts prior to polymerization), post-processing techniques (Chemical/thermal exfoliation, defect etching, and ultrasonication), structure–property relationships (how SSA enhancement influences band structure, carrier dynamics, and catalytic efficiency), and multifunctional applications (environmental remediation, energy conversion, and emerging biomedical uses (Figure 1)). Finally, we critically analyze remaining challenges (e.g., scalability, stability) and propose future research directions to guide the rational design of high-performance g-C3N4-based materials with optimized surface architectures.

2. The Method of Synthesizing High-SSA Carbon Nitride

The aforementioned analysis of application trends reveals that the performance of g-C3N4 in energy conversion and environmental remediation is critically dependent on its surface/interface properties. However, bulk g-C3N4 prepared by conventional methods suffers from low specific surface area (<20 m2/g) and insufficient active site exposure, which fundamentally limit its photocatalytic/electrocatalytic efficiency. To address these bottlenecks, recent research has focused on controlled synthesis strategies at the molecular level, including topological structure design, defect engineering, and heterojunction construction, to precisely tailor the specific surface area and surface chemistry of g-C3N4. The following section systematically explores synthetic paradigms for high SSA g-C3N4, elucidating how these strategies optimize mass transport pathways, enhance light harvesting, and promote charge carrier separation, thereby providing a material foundation to overcome the application challenges discussed above.

2.1. Structure of Carbon Nitride

In the structure of g-C3N4, the C and N atoms are both sp2 hybridized and their p orbitals have overlapped to form delocalized Π-bond. The basic units are connected to each other by N atoms at the end to form a planar grid structure; g-C3N4 then presents a layer stacking structure, which is similar to that of graphite (Figure 2). The DFT study of g-C3N4 illustrates that it has the special properties of a semiconductor, such as a suitable band gap (~2.6 eV), and the appropriate positions of valence band and conduction band. These results indicate that g-C3N4 has a certain redox capacity in theory (oxidizing water to produce oxygen, reducing water to produce hydrogen), which lays the foundation for its application in the field of photocatalytic hydrogen production.

2.2. Preparation of Carbon Nitride

g-C3N4 is usually formed by a precursor that contains C and N elements, through thermal (550–600 °C) polymerization, which represents an attractive method due to its simplicity and low cost. Cyanamide, dicyandiamide, melamine, urea, and thiourea commonly serve as N-rich precursors. In the case of melamine as the precursor, a 3-s-triazine ring can be formed after the polymerization and rearrangement of melamine. Condensation into g-C3N4 then continues with the 3-s-triazine ring as the basic unit. NH3 is released during the polymerization process. Therefore, the structural properties of g-C3N4 strongly depend on the synthesis conditions.

2.3. Preparation of High-SSA Carbon Nitride

In addition, researchers frequently employ templating methods, precursor selection and processing, pyrolysis conditions, among others, to prepare g-C3N4 with large SSA. In recent years, solvothermal methods and supramolecular assembly techniques have also gained prominence. Here, we present a statistical overview of these various approaches, categorizing them comprehensively, and detailing the intricacies, advantages, and limitations of each method. The preparation methods of g-C3N4 with large SSA mainly focus on the following three aspects. High-SSA g-C3N4 with a special structure has been prepared by changing the thermal polymerization mode of the N-rich precursors through special pretreatment. Processing during the synthesis of g-C3N4 by introducing a template and salt reagent can prepare g-C3N4 with a special morphology. Post-processing of the synthesized g-C3N4 by adding energy and acid/base reagents can be used to destroy the interlayer forces and produce pore structures. This critical review presents a hierarchical analysis framework for synthesizing high-SSA g-C3N4 materials through a tripartite approach: precursor engineering (pre-treatment), in-process treatment, and post-synthesis. The established structure–property correlation matrix enables rational selection of synthesis protocols based on application-specific performance metrics.

2.3.1. Pre-Treatment of Carbon Nitride

The Influence of Raw Materials on the Specific Surface Area and Morphology
The selection and mix of raw materials are part of the simple, template-less, one-step polymerization method for the preparation of g-C3N4 with high SSA (Table 1) [20,21,25,26,27,28,29]. When urea is used as a precursor, CO2 is released simultaneously, resulting in the synthesized g-C3N4 showing a more porous structure and larger SSA. Additionally, it is worth noting that the mixed raw materials can be assembled into a supramolecular precursor by hydrogen bonding in solution [30]. The structure of the prepared g-C3N4 ultimately depends on the supramolecular precursor, because the melting point of supramolecular precursor is higher than the formation temperature of bulk g-C3N4 during thermal polymerization (Figure 3). Wang et al. prepared g-C3N4 nanotubes through thermal polymerization by mixing urea with melamine (mass ratio = 10:1) [28]. In thermal polymerization, urea can be converted to cyanuric acid at 400 °C, which then combined with melamine by hydrogen bonding to form supramolecular nanorods. Therefore, the formation of supramolecular precursors is closely related to the mass ratio of urea and melamine and further guides the formation of nanotubes (Figure 4a). Chen et al. prepared 3D g-C3N4 with a SSA of 130 m2 g−1 by mixing melamine and cyanic acid (at a molar ratio of 1:1) [29]. The 3D interconnected open-framework provided a carrier transport channel for highly efficient photocatalytic overall water splitting (splitting pure water into H2 and O2 with high evolution rate up to 101.4 and 49.1 µmol g−1 h−1). Bao et al. designed a simple template-mediated approach by supramolecular self-assembly (Cu-melamine-cyanuric acid) to prepare a copper-doped porous g-C3N4 (Cu-pCN) photocatalyst, increasing the SSA from 11.37 to 142.8 m2 g−1 [31].
The Treatment for the Precursor of Carbon Nitride
The synthesis of g-C3N4 with special nanostructure is undoubtedly one of the effective ways to improve its specific surface area. As the performance of g-C3N4 depends on its nitrogen precursor, some treatment methods for the precursor have gradually attracted wide attention. Apart from the precursor mixing, treatment for the precursor of g-C3N4 has attracted increasing attention because the properties of g-C3N4 are closely associated with the nitrogen-rich precursor (Figure 3). This method means that the precursor is treated with acid [32,33,34,35], organic solvent [36,37], and hydrothermal process [38,39,40,41] before the normal calcination process. Through pretreatment, the precursor can form a more abundantly porous structure during the subsequent thermal polymerization or solvothermal reaction, thereby significantly increasing the specific surface area of the g-C3N4. The enlargement of the specific surface area implies more active sites for catalytic reaction, which contributes to enhancing the catalytic performance of g-C3N4. Additionally, pretreatment can also assist in controlling the morphology of g-C3N4, such as achieving a sheet-like or porous plate-like structure. These specific morphologies facilitate the migration of photogenerated carriers and the mass transfer of reactants, further enhancing the material’s photocatalytic performance.
Dong et al. prepared g-C3N4 with a porous ultrathin nanosheet structure by calcining HCl-treated melamine [32]. The thermal polymerization mode of HCl-treated melamine is different from the traditional mode, due to the HCl reacting with the amino groups of melamine. The production of voids increased the specific surface area (345 m2 g−1) and reduced the charge migration distance (from body to surface), resulting in an excellent NO photocatalytic removal rate. Zhang et al. speculated that acid could change the growth orientation of the precursor, which would be beneficial to the increase of active sites [33]. Additionally, a defect-rich amorphous g-C3N4 photocatalyst (228.4 m2 g−1) was synthesized by simple direct calcination of the rationally size-reduced urea crystals, which was effectively controlled by the anti-solvent growth method [36] (Figure 4b). The introduction of N vacancies caused a broad visible-light-responsive range, exposed surface bonding sites, and quenched radiative recombination, leading to enhanced photocatalytic activity for hydrogen production (37,680 µmol g−1 h−1 under visible-light irradiation). Additionally, hydrothermal processes have also been used to change the precursor, affecting the structure of g-C3N4. The hydrolytic product of the precursor can assemble with the precursor to form a supramolecular precursor. Mo et al. synthesized g-C3N4 nanotubes with a specific surface area of 127.8 m2 g−1 by calcining the hydrothermal treated melamine [38]. Compared with bulk g-C3N4, the g-C3N4 nanotubes exhibited an excellent hydrogen evolution rate (118.5 µmol h−1), which was ascribed to the nanotube structure and the N defects. Moreover, Cheng et al. suggested that the hydrolytic products of dicyandiamide (amidine urea) can weaken the interaction between Π–Π stacking, which is helpful for generating g-C3N4 in thin layers [40]. Therefore, the precursor pretreatment can not only change the morphology but also optimize the electronic structure of g-C3N4. However, the performances of the obtained g-C3N4 are closely related to the details of the treatment, such as the concentration of acid, organic solvent, and the temperature and duration of the hydrothermal process (Table 2) [32,33,34,35,36,37,38,39,40,42,43]. In order to achieve high-performance g-C3N4, the optimization of details in the pretreatment process needs to be further considered.
Figure 3. Pre-treatment of high-SSA g-C3N4 [20,25,28,36,38,43].
Figure 3. Pre-treatment of high-SSA g-C3N4 [20,25,28,36,38,43].
Nanomaterials 15 00956 g003
Figure 4. (a) The formation process of hollow C3N4 nanotubes. Reproduced with permission [28]. Copyright © 2019 Springer Nature. (b) Preparation process and photocatalytic activity of defect-rich amorphous carbon nitride; hydrogen-evolution rate for defect-rich amorphous carbon nitride in comparison with other GCN-based photocatalysts. Reproduced with permission [36]. Copyright © 2018 American Chemical Society.
Figure 4. (a) The formation process of hollow C3N4 nanotubes. Reproduced with permission [28]. Copyright © 2019 Springer Nature. (b) Preparation process and photocatalytic activity of defect-rich amorphous carbon nitride; hydrogen-evolution rate for defect-rich amorphous carbon nitride in comparison with other GCN-based photocatalysts. Reproduced with permission [36]. Copyright © 2018 American Chemical Society.
Nanomaterials 15 00956 g004

2.3.2. In-Process Preparation of Carbon Nitride

Introducing the Template During the Synthesis
Template methods are widely used for precise control of the morphology and porous structure of materials, according to a variety of templates. The prepared g-C3N4 by template method usually displays a mesoporous structure and a large specific surface area. Common templates for preparing g-C3N4 include SBA-15, MCM-22, SiO2, KCC-1 silica spheres (Table 3) [44,45,46,47,48,49,50,51,52,53,54,55,56,57,58]. SBA-15 has been used as a template to synthesize g-C3N4 with uniform mesoporous structure, showing a high specific surface area of 505 m2 g−1 [44]. Cui et al. prepared g-C3N4 with specific surface areas of 188 m2 g−1 and 239 m2 g−1 using SiO2 and SBA-15, respectively [47]. This shows that the morphologies and specific surface areas of the prepared g-C3N4 depend on the template used. In addition, Vinu et al. synthesized mesoporous g-C3N4 (239 m2 g−1) with different pore sizes, using SBA-15 [45]. It should be noted that the pore volume, pore size, specific surface area, and N content of the prepared g-C3N4 can be controlled by the simple adjustment of the ethylenediamine–carbon tetrachloride weight ratio. Therefore, the structural properties of the prepared g-C3N4 are also related to the precursor; the well mixing between precursor and template is one of the important steps of the template method. For example, Zhao et al. prepared ordered mesoporous g-C3N4 nanorods by using hexamethylenetetramine as the precursor, which had high specific surface area (1116 m2 g−1), a bimodal mesoporous structure, and high N content [48]. Kailasam et al. mixed cyanamide (precursor for g-C3N4) directly with tetraethyl silicate (precursor for silica) using the sol–gel method to ensure the simultaneous production of g-C3N4 and SiO2 and achieved full mixing [51].
Besides that, the g-C3N4 with special nanostructures have been prepared by making various special templates. Inspired by natural photosynthesis, Tong et al. prepared the multiple-shell g-C3N4 (310.7 m2 g−1) with excellent photocatalytic hydrogen production activity, imitating the excellent light capture ability and water storage structure of chloroplasts (Figure 5a) [53]. The successful preparation of this structure depended on the self-made three-layer spherical SiO2 template with a porous structure. Specifically, the existence of pores in the SiO2 template ensured the full mixing between cyanamide and template. Lin et al. synthesized mesoporous g-C3N4 (132.26 m2 g−1) with a chiral supramolecular helical structure and liquid crystal layered structure via prepared chiral mesoporous SiO2, exhibiting an ultrahigh hydrogen evolution rate of 219.9 µmol h−1 under visible-light irradiation compared with bulk g-C3N4 [54]. Liu et al. mixed mesoporous SiO2 nanorods and cyanamide several times to prepare mesoporous g-C3N4 nanotubes with a specific surface area of 135.1 m2 g−1 using a nano-confined reaction inside silica nanotubes with porous shells [55]. Moreover, dicyandiamide and melamine can also be used in the preparation of g-C3N4 with a high specific surface area by changing the mixing mode of the template and precursor. For example, onion-like ring g-C3N4 was prepared using melamine and SiO2 [56]. The vapor deposition method was selected for the synthesis, because solid melamine is easy to sublimate in the pyrolysis process. Si et al. synthesized hierarchical porous PCN microspheres (Hydrogen evolution activity of 4635.8 µmol h−1 g−1) via in situ precursor conversion technology utilizing the solubility difference between cyanamide and dicyandiamide in aqueous solution [57]. After the addition of NH4OH, the precursor started to transform from cyanamide to dicyandiamide along the wall surfaces of the hierarchical porous SiO2 microspheres. This procedure can overcome the low solubility of dicyandiamide and avoid the evaporation of cyanamide.
Introducing the Salt During the Synthesis
Moreover, the polymerization process of the precursor can be affected by salt [15,42,43,59]. Salts with low melting points gradually melt to form a liquid phase under high-temperature conditions. The raw materials undergo reactions in the liquid phase of the molten salt, which not only facilitates the mass transfer and diffusion processes of the reactant particles but also benefits crystal growth. Additionally, some molten salt particles penetrate into the gaps between the particles of the generated product, preventing the aggregation of product particles and improving the crystal structure of the sintered product. For example, Tian et al. controlled the synthesis of g-C3N4 hollow tubes with a specific surface area of 128 m2 g−1 through the molten salt method [43]. First, the melamine was oversaturated and nucleated in advance by adding a LiCl/KCl eutectic mixture during the polymerization process. Then, the formed nanoparticles combined and grew into nanosheets in the presence of highly soluble Li+, K+, and Cl. Finally, the nanosheets curled to form the g-C3N4 hollow tube due to the minimization of surface free energy (Figure 5b). In addition, N species could easily be captured in the living salt medium, leading to a high N content in the synthesized g-C3N4.

2.3.3. Post-Processing of Carbon Nitride

Similar to the reason why it is difficult to exfoliate layers of graphene, the layers of g-C3N4 agglomerate due to the presence of van der Waals forces, resulting in a relatively low specific surface area. Inspired by the Hummers method of preparing graphene, a series of post-processing methods for exfoliating g-C3N4 have been developed due to its graphite-like layered stacking structure. g-C3N4 can be intercalated by acids and organic solvents. The interlayer forces are destroyed by the introduced energy (ultrasonic, thermal energy, high temperature, and pressure) to form g-C3N4 nanosheets. In addition, the addition of base can cause reactions with C or N to form porous structures. All the above approaches can increase the specific surface area of g-C3N4 (Figure 6). Recently, each method has been combined with other methods to exfoliate g-C3N4 in a more convenient, energy-saving, and efficient was. The large surface area of the g-C3N4 nanosheets that are produced can increase the area of contact with reactants. Moreover, the nanosheets’ structure shortens the migration path of photo-generated carriers from the catalyst’s bulk to its surface, which effectively inhibits the recombination of photo-generated carriers during the transfer process.
Acid/Chemical Treatment
In an acidic solution, hydrogen ions (H+) undergo a protonation reaction with the nitrogen atoms on the surface of g-C3N4 or other basic sites, significantly increasing the electrostatic repulsion between the layers, effectively weakening the van der Waals forces between the layers and reducing the bonding strength between them. Meanwhile, the treatment of g-C3N4 in liquid acid is a means of non-oxidized intercalation, exfoliation, and surface modification, which can preserve the C-N heterocyclic skeleton and avoid the excessive production of defects [14,60,61]. Numerous works have combined acid with ultrasonication or thermal treatment to increase the specific surface area of g-C3N4 efficiently [62,63,64,65].
For example, g-C3N4 ultra-thin two-dimensional nanosheets with a micro-mesoporous structure were prepared by inserting H3PO4 into g-C3N4 layers [60]. The obtained g-C3N4 showed a larger specific surface area (increasing from 7.6 m2 g−1 to 55.4 m2 g−1) and more transverse charge transfer channel to promote the separation of the photo-generated carriers (Figure 7a). The average H2 evolution rate under visible light is 195.8 µmol h−1, which is much higher than that of bulk g-C3N4 (14.5 µmol h−1). Du et al. synthesized large-scale soluble acidified g-C3N4 (42 m2 g−1) via H2SO4 acidification, expanding the functionalization and application of g-C3N4 [61]. Bulk g-C3N4 was exfoliated using a combination of H2SO4 and ultrasonication to form nanosheets (25.7 m2 g−1); with a higher reaction rate of p-nitrophenol photo-reduction due to the positively charged surface and the nanosheets’ morphology [62]. In the same way, monolayer C3N4 nanosheets with a specific surface area of 205.8 m2 g−1 were obtained through sonicated exfoliation after H2SO4 treatment, showing a threefold enhancement in photocatalytic H2 production [65]. Ma et al. selected HCl in combination with sonication; the as-prepared ultrathin g-C3N4 nanosheets had a high specific surface area of 305 m2 g−1 and reached the lowest heparin detection limit of 18 ng mL−1 [63]. Thus, it can be seen that the exfoliation effect on g-C3N4 is closely related to the type and concentration of the liquid acid, the mixing time, and the follow-up details (such as speed at subsequent centrifugation). The application of this method is restricted by uneven treatment, multiple influencing factors, long processing time, and low yield. In addition, g-C3N4 nanosheets prepared via liquid acid exfoliation showed a wider band gap due to the quantum size effect, resulting in diminished visible-light response (Table 4) [60,61,62,63,64,65,66,67].
In addition, the Hummers method had been used to exfoliate bulk g-C3N4, due to its graphite-like structure [66,67]. Teng et al. synthesized ultrathin g-C3N4 nanosheets using combination of intercalation with H2SO4, oxidant-assisted exfoliation of KMnO4, and reduction of HHA [66]. The obtained g-C3N4 had a large specific surface area of 355.8 m2 g−1, a wider band gap of 3.07 eV, and an increased lifetime. However, the structure of g-C3N4 may be destroyed by the use of strong oxidants, resulting in collapse of the skeleton or an increase in defects.
Table 4. Influencing factors, assistant methods, and results of adding acids/chemical exfoliation [60,61,62,63,64,65,66,67].
Table 4. Influencing factors, assistant methods, and results of adding acids/chemical exfoliation [60,61,62,63,64,65,66,67].
ReferencesSynthesis of Bulk g-C3N4AcidAssistant MethodResults
Mbulk g-C3N4TypeConcentration or VolumeTreated TimeBand GapBETMorphologyApplication
[60]Dicyandiamide 550 °C for 4 h at 5 °C min−10.2 gH3PO40.3 mL3 dDispersed in 20 mL dimethylformamide2.63 eV

2.74 eV
7.6 m2/g

55.4 m2/g
Nanomaterials 15 00956 i034Photocatalytic hydrogen evolution and CO2 conversion
[61]Dicyandiamide 550 °C for 4 h at 2.3 °C min−14 gH2SO4 (98%)52 g5 h20 g oleum
NH4Cl (1.60 mol)
2.83 eV

3.78 eV
12 m2/g

42 m2/g
Nanomaterials 15 00956 i035-
[62]Melamine2 gH2SO440 mL 98 wt%10 hSonicated with 8 h2.64 eV

2.68 eV
8.1 m2/g

25.7 m2/g
Nanomaterials 15 00956 i036Photo-reduction of p-nitrophenol
[63]Dicyandiamide 550 °C for 4 h at 2.3 °C min−1 in static air1 gHCl25 mL 10 M1 hSonication for 2 h2.57 eV

2.75 eV
9 m2/g

305 m2/g
Nanomaterials 15 00956 i037Excellent metal-/label-free biosensing platform
[64]Melamine 550 °C for 4 h at 3 °C min−1 in static air1.0 gH2SO410 mL 98 wt%8 hCalcinated at 550 °C in N2
for 2 h
2.65 eV

2.79 eV
12 m2/g

54.3 m2/g
Nanomaterials 15 00956 i038Photocatalytic degradation
[65]Dicyandiamide 550 °C for 4 h at 2.3 °C min−1 in static air1 gH2SO410 mL 98 wt%8 hSonication2.64 eV

2.92 eV
4.3 m2/g

205.8 m2/g
Nanomaterials 15 00956 i039Photocatalytic hydrogen evolution
[66]-1 gH2SO430 mL 98 wt%-Hummer’s method2.54 eV

3.07 eV
3.5 m2/g

355.8 m2/g
Nanomaterials 15 00956 i040-
[67]Dicyandiamide 500 °C for 1 h 520 °C for 3 h10 gH2SO4230 mL 98 wt%-Hummer’s method2.72 eV

3.85 eV
-Nanomaterials 15 00956 i041Photocatalytic degradation
Figure 6. Pre-treatment of high-SSA g-C3N4 [60,63,65,68,69,70,71,72,73,74,75,76].
Figure 6. Pre-treatment of high-SSA g-C3N4 [60,63,65,68,69,70,71,72,73,74,75,76].
Nanomaterials 15 00956 g006
Figure 7. (a) Illustration of the preparation process, SEM images, photocatalytic activity for H2 evolution, and photocatalytic H2 evolution stability of P-PCNNS under visible light. Reproduced with permission [60]. Copyright © 2016 WILEY-VCH. (b) Schematic illustration of liquid-exfoliation process, suspension, and crystal structure of ultrathin C3N4 nanosheets, viability after 48 h, confocal fluorescence image, and overlaid bright-field image of HeLa cells incubated with ultrathin C3N4 nanosheets. Reproduced with permission [68]. Copyright © 2012 American Chemical Society. (c1) Schematic of the formation process of C3N4 nanosheets and hydrogen evolution from water under UV-visible light. Reproduced with permission [69]. Copyright © 2012 WILEY-VCH. (c2) Top-down process for preparation and cumulative surface area of foam-like hollow ultrathin C3N4 nanosheets. Reproduced with permission [77]. Copyright © 2016 WILEY-VCH. (d) The intralayer chemical structure of C3N4 and the amount of H2 versus time after hydrothermal treatment. Reproduced with permission [70]. Copyright © 2018 Elsevier.
Figure 7. (a) Illustration of the preparation process, SEM images, photocatalytic activity for H2 evolution, and photocatalytic H2 evolution stability of P-PCNNS under visible light. Reproduced with permission [60]. Copyright © 2016 WILEY-VCH. (b) Schematic illustration of liquid-exfoliation process, suspension, and crystal structure of ultrathin C3N4 nanosheets, viability after 48 h, confocal fluorescence image, and overlaid bright-field image of HeLa cells incubated with ultrathin C3N4 nanosheets. Reproduced with permission [68]. Copyright © 2012 American Chemical Society. (c1) Schematic of the formation process of C3N4 nanosheets and hydrogen evolution from water under UV-visible light. Reproduced with permission [69]. Copyright © 2012 WILEY-VCH. (c2) Top-down process for preparation and cumulative surface area of foam-like hollow ultrathin C3N4 nanosheets. Reproduced with permission [77]. Copyright © 2016 WILEY-VCH. (d) The intralayer chemical structure of C3N4 and the amount of H2 versus time after hydrothermal treatment. Reproduced with permission [70]. Copyright © 2018 Elsevier.
Nanomaterials 15 00956 g007
Ultrasonication
Liquid-phase ultrasound has attracted much attention due to its simple and mild exfoliation process, which can exfoliate bulk g-C3N4 to achieve a nanosheet structure [68,71,72,78]. The matching degree of the surface energy between the used solvent and g-C3N4 determines effective exfoliation via liquid-phase ultrasound. The relationship between the surface energy and mixing enthalpy can be described by the following empirical formula [71]:
Δ H m i x V m i x = 2 T b u l k ( δ G δ s o l ) 2 ϕ
where ΔH is the enthalpy of mixing, δ is the square root of the component surface energy, Tbulk is the average thickness of g-C3N4, and ϕ is the volume fraction of g-C3N4. It is well known that the bulk g-C3N4 can be effectively exfoliated by liquid-phase ultrasound when the mixing enthalpy is minimized; that is, when the surface energy of g-C3N4 and the solvent match.
Liquid-phase ultrasound is usually selected in combination with other methods to exfoliate bulk g-C3N4 in order to increase its specific surface area. For example, ultrathin g-C3N4 nanosheets have been successfully prepared in water via green liquid phase exfoliation, and it has been used in biological imaging (Figure 7b) [68]. In detail, the bulk g-C3N4 was treated ultrasonically in water (16 h) to obtain atom-thick nanosheets. The prepared g-C3N4 nanosheets showed enhanced capacity for light absorption and response, inducing an extremely high PL quantum yield, beneficial for bioimaging application. Meanwhile, organic solvents have also been selected as dispersion media for liquid-phase ultrasound due to their own functional groups [71,72,78]. Yang et al. synthesized g-C3N4 nanosheets by liquid-phase ultrasound in IPA [71], with a high specific surface area (384 m2 g−1) and a large aspect ratio, beneficial for photocatalytic hydrogen evolution. She et al. compared the exfoliation effect between butanediol and 1.3-BUT on bulk g-C3N4, demonstrating the influence of functional groups on the exfoliation process [78]. The existence of two hydroxyl groups (1.3-BUT) leads to more effective exfoliation of g-C3N4, resulting in the enhanced microelement-sensing capability. Moreover, Lin et al. conducted ultrasonic treatment of g-C3N4 with mixed solvent as the dispersing medium [72]. In the aqueous solution with adjustable concentration, g-C3N4 nanosheets (59.4 m2 g−1) were prepared by changing the volume ratio of the two solvents. Thus, the properties of the organic solvent are closely related to the exfoliation effect of the liquid-phase ultrasound.
In addition, the interlayer forces and surface properties of g-C3N4 can be influenced by its crystallization and polymerization. Thus, the exfoliation effect of liquid-phase ultrasonic treatment is also related to the method of synthesis of the g-C3N4. The application of the liquid-phase ultrasonic method has been limited due to its long processing time and unsatisfactory yield (Table 5) [68,71,72,78]. Recently, liquid-phase ultrasonic has gradually become an auxiliary method to prepare high-surface-area g-C3N4.
Thermal Oxidation Treatment
It is well known that g-C3N4 has good thermal stability and can maintain its basic structure at medium-high temperatures (< 700 °C). However, the stability between g-C3N4 layers is reduced at medium-high temperature, which can cause destruction of the force between the layers. Thus, thermal oxidation treatment has been widely applied for the exfoliation of bulk g-C3N4. Moreover, the required processing time for thermal oxidation treatment to prepare high-surface-area g-C3N4 is shorter than that for liquid-phase ultrasound [11,23,69,79]. For example, g-C3N4 nanosheets with a specific surface area of 306 m2 g−1 and a thickness of 2 nm were prepared by thermal oxidation treatment of bulk g-C3N4 in air (at 500 °C for 2 h with 5 °C min−1) [69]. The average hydrogen evolution rate of the g-C3N4 nanosheets under UV-visible light was 170.5 μmol h−1, which is 5.4 times higher than that of the bulk g-C3N4. This can be ascribed to the synergistic effects of large surface area, increased bandgap, improved electron transport ability, and prolonged lifetime of the charge carriers. Additionally, the quality of the obtained material is inversely proportional to the processing time (Figure 7(c1)), because part of the g-C3N4 is oxidized to produce gas. In order to facilitate the exfoliation of g-C3N4, numerous studies have attempted to improve thermal oxidation treatment methods. Inspired by the preparation of highly corrugated graphene sheets, the high-surface-area g-C3N4 has been obtained by combining a rapid cooling process with thermal oxidation treatment [73,80]. Zhang et al. prepared g-C3N4 nanosheets (60.51 m2 g−1) via high-temperature treatment and a subsequent cooling process [80], heating the bulk g-C3N4 at 800 °C for 15 min with a heating rate of 600 °C min−1, then putting it in water to cool (15 °C). In our previous work, g-C3N4 nanosheets (142.8 m2 g−1) were prepared by rapid cooling after the heating process [73]. During the process of formation, the inter-layer force was broken by the shrinkage force generated from the rapid and large temperature difference, causing the gradual exfoliation of the g-C3N4. In addition, Zhang et al. synthesized g-C3N4 nanosheets (117.27 m2 g−1) by a continuous heating process, preparing the nanosheet structure with a smaller pore size by extending the thermal treatment time [74].
However, the degree of exfoliation in a single batch of samples may be different due to uneven heating in the thermal oxidation treatment. Therefore, bulk g-C3N4 is usually treated in small quantities to ensure uniform heating, resulting in a low yield. To solve this, Zhao et al. prepared g-C3N4 nanosheets with a high specific surface area (increasing from 10.94 m2 g−1 to 99.73 m2 g−1) by combining thermal oxidation treatment with liquid-phase ultrasound [81]. Reheating the dispersed g-C3N4 treated by ultrasound can avoid the occurrence of uneven heating, thus improving the yield of nanosheets. In addition to low yield, the application of thermal oxidation treatment for exfoliating bulk g-C3N4 has been limited by the single morphology of prepared g-C3N4 nanosheets. Recently, numerous works have prepared nanosheets with porous structures, including increased duration of treatment [77,82], changes of atmosphere [83,84,85], and addition of reagents [13,86,87,88,89]. For example, Li et al. extended the heating time to 6 h for preparing ultrathin g-C3N4 nanosheets (277.98 m2 g−1) with abundant micropores and mesopores [77]. Firstly, the interlayer forces have been broken by thermal oxidized process, then some basic units have been etched by increasing the calcination time, leading the formation of nanosheets with porous structures (Figure 7(c2)). The presence of abundant porous structures exposes more active edges and provides a cross-surface diffusion channel. It can promote the separation and transfer of photo-carriers, which is beneficial to the improvement of photocatalytic activity. Huangfu et al. prepared ultrathin g-C3N4 by repeated use of this simple heating method [82]. In addition, Hou et al. prepared porous nanosheet g-C3N4 with a specific surface area of 114 m2 g−1 by injecting hydrogen during the heating treatment [83]. There were many N vacancies in the structure, which overcame the quantum size effect, resulting in the simultaneous small size and narrow band gap. g-C3N4 with a porous structure (196 m2 g−1) and C vacancies was prepared by thermal oxidation treatment in ammonia gas, which enhanced the capacities of mass transfer, charge transfer, and light absorption [84]. Furthermore, Song et al. prepared the porous g-C3N4 nanosheets (265.2 m2 g−1) by thoroughly mixing bulk g-C3N4 with potassium hydroxide before reheating treatment in the air [86]. In this process, the adding alkali reacted with the C atoms to produce a porous structure for accelerating mass transfer, charge transfer, and the exposure of active sites. Hu et al. synthesized bimodal porous g-C3N4 nanosheets with a specific surface area of 219 m2 g−1 by the same method, which exhibited an improved photocatalytic hydrogen evolution of 1900 µmol h−1 g−1 (8.6 times higher than that of bulk g-C3N4) [87].
Although thermal oxidation treatment can exfoliate the bulk g-C3N4 effectively, low yields still restrict its further application. Due to the decomposition of C3N4 at high temperature, the yield of obtained g-C3N4 decreases with an increase in heating time and temperature. Moreover, the effect of exfoliation is closely related to the synthesis details in the process, including temperature, heating speed, treatment time, dosage, and atmosphere (Table 6) [69,73,74,77,79,80,81,82,83,84,85,86,87]. Thus, precise control of the trade-off between specific surface area and yield of exfoliated g-C3N4 still needs to be developed. Recently, thermal oxidation treatment has been combined with other methods to prepare g-C3N4 with a larger surface area.
Post-Hydrothermal
As one of the common methods used in preparation and treatment of materials, the hydrothermal method (solvothermal method) has also been applied in the post-processing of g-C3N4. Bulk g-C3N4 is usually placed in solution for subsequent hydrothermal treatment [12,70,75,76,90]. During this process, the presence of alkali can promote the formation of a porous structure and the surface modification of g-C3N4. Nie et al. prepared g-C3N4 with a fluffy porous structure (64.7 m2 g−1) via an alkali hydrothermal method [75]. The NO photocatalytic removal activity can be attributed to the increased specific surface area, narrow band gap, and low carrier recombination rate. However, the porous structure can be excessively enlarged or even destroyed by high concentrations of sodium hydroxide. In addition, ammonium hydroxide has been used in hydrothermal treatment of bulk g-C3N4 [70]. The C atoms can be partial oxidized by ammonium hydroxide, causing the formation of -OH and -C=O- on the surface of g-C3N4, which can promote the interaction of photocatalyst and reactant (Figure 7d). Moreover, the substitution of N is beneficial to the redox ability on the surface reaction sites. However, the excessive use of ammonium hydroxide can also cause structural damage, increasing the presence of sites for the recombination of photo-generated carriers. Wang et al. speculated that the proton exchange between -OH and -NH/NH2 affected the electron properties when the alkali concentration was too high and then inhibited the photocatalytic reaction [90]. Obviously, the main roles of hydrothermal method in treating bulk g-C3N4 are to promote the formation of the pore structure and surface modification. The concentration of alkali in the hydrothermal process needs to be explored to avoid excessive destruction of the g-C3N4 structure (Table 7) [70,75,76,90,91,92].
Recently, this method has been combined with other treatments to exfoliate bulk g-C3N4, which can make up for the drawbacks of the hydrothermal process, such as excessive defects and small specific surface area. For example, Zeng et al. prepared sea-urchin-structure g-C3N4 by combining sealed condensation with hydrothermal treatment [91]. The porous g-C3N4 was prepared by partial decomposition of g-C3N4 during the sealed condensation. Subsequently, this particular structure with a specific surface area of 65.6 m2 g−1 was formed by hydrothermal treatment. The sea-urchin-structure g-C3N4 showed a narrow band gap of 2.0 eV, overcoming the quantum size effect. Most importantly, its conduction band and valence band potential were suitable to promote overall water splitting under visible light irradiation. Additionally, the alkali hydrothermal process can be combined with carbon thermal reduction to optimize the morphology and electronic configuration of g-C3N4 [92]. The surface hydroxylation and exfoliation of g-C3N4 were performed via the former, and the electron configuration was optimized via carbon thermal reduction. The prepared g-C3N4 (197 m2 g−1) had a remarkably improved photocatalytic hydrogen evolution rate of 246.2 μmol h−1 under visible light irradiation (λ > 420 nm), 17-fold higher than that of bulk g-C3N4.
The above post-processing methods used in exfoliation of g-C3N4 are still restricted by the excessive defects (chemical exfoliation), long processing times (ultrasonication), low yield (thermal oxidation treatment), and structural damage (post-hydrothermal method). Gradually, the combination of two different methods has become a common pattern for the exfoliation of bulk g-C3N4 to make up for the shortcomings of single exfoliation methods. However, the details of the combinations between these methods need to be further explored. Furthermore, the entire design between band structure and morphological control of g-C3N4 should be considered to realize the excellent photocatalytic ability of g-C3N4.

2.4. Summary

In summary, precursor mixing has emerged as a versatile and sustainable strategy for template-free synthesis of morphologically diverse g-C3N4 architectures with enhanced SSA. This approach circumvents the need for energy-intensive processes, toxic template removal, or additional reagents, offering inherent advantages in cost-effectiveness and environmental compatibility (Figure 8). As evidenced by the steadily rising publication trend in related studies from 2016 to 2023 (Figure 9), this methodology has garnered significant attention as a mainstream paradigm for engineering g-C3N4 with tailored nanostructures.
The mechanistic foundation of this strategy lies in the supramolecular preorganization of precursor complexes, where the selection and combinatorial principles of nitrogen-rich precursors dictate the self-assembly pathways. As illustrated in Figure 10, commonly employed precursors include melamine–cyanuric acid systems, urea–thiourea hybrids, doped precursors like Cu–melamine complexes; typical mixing modalities involve solvent-mediated hydrogen bonding, stoichiometry-controlled co-crystallization, and pH-triggered electrostatic assembly.
Crucially, the interplay between precursor chemistry (e.g., hydrogen-bonding capacity, π–π stacking propensity) and processing parameters (molar ratios, solvent polarity) governs the formation of intermediate supramolecular architectures, which template the final g-C3N4 morphology during thermal condensation. For instance, melamine–urea systems achieve nanotube formation through temperature-dependent cyanurate intermediate assembly, while acid-mediated precursor protonation induces lamellar exfoliation. Systematic mapping of these relationships, through combinatorial precursor screening coupled with machine learning-driven optimization, could enable predictive design of g-C3N4 catalysts with application-optimized SSA and defect landscapes.

3. Research Distribution and Trend Analysis of Carbon Nitride Applications

The escalating environmental pollution and energy crises demand transformative solutions, driving the exploration of advanced semiconductor materials. Among them, g-C3N4 has emerged as a versatile platform due to its tunable electronic structure, visible-light responsiveness, and earth-abundant composition. Our bibliometric analysis (Figure 11a) reveals a declining trajectory in g-C3N4 research across traditional domains (2016–2023), with energy and environmental applications dominating ~80% (decreasing from 90%-2019 to 80%-2023) of publications (Figure 11b). Notably, emerging trends highlight a shift to in environmental field from dye degradation (60% decline since 2016) to emerging refractory pollutants (antibiotics: +40%) (Figure 12a), a transition in the energy sector from photocatalytic H2 production (−30% since 2016) to CO2 reduction (+20%) (Figure 12b), and steady growth in antibacterial and biosensing research for biomedical applications (~10% of total) (Figure 11b). This evolution reflects the material’s adaptability to address sustainability imperatives and technological bottlenecks. The following sections dissect these trends with mechanistic insights and case studies.

3.1. Environmental Remediation: From Conventional Pollutant Removal to Emerging Contaminants

g-C3N4-based systems excel in multiphase pollutant management through light-driven redox cycles (Table 8) [21,32,33,37,39,43,49,62,73,79,83,93,94,95,96,97]. The photocatalytic mechanism involves photon absorption (λ > 420 nm for visible-light activation), carrier generation and migration (e-h+ pairs with CB~−1.3 V vs. NHE), formation of reactive species (•OH, •O2 via H2O/O2 activation), and pollutant mineralization (e.g., aromatic ring cleavage in antibiotics). Furthermore, a high SSA directly enhances remediation efficacy through the enlarged pollutant adsorption capacity and the increased density of active sites for redox reactions.

3.1.1. Aquatic Organic Contaminant Degradation

Dyes, nitrogen-containing organic compounds, phenols, and antibiotics are often used as target pollutants to test the degradation ability of g-C3N4 materials [85,98,99,100,101]. While early studies focused on model organic dyes (e.g., methylene blue), recent efforts have targeted priority pollutants. For example, the photodegradation activity of polyaniline (PANI)/carbon nitride nanosheet (CNNS) composite hydrogel was tested via the degradation of methylene blue. The excellent performance in removing organic pollutant can be ascribed to the cooperation of adsorptive preconcentration and the subsequent photocatalytic oxidation [98]. Zhang et al. constructed a Bi7O9I3/C3N4 Z-scheme heterojunction photocatalyst for degrading doxycycline hydrochloride under visible light [99]. The dominant oxhydryl and superoxide active groups led to excellent photodegradation and mineralization ability for doxycycline hydrochloride. Moreover, the prepared tetragonal carbon nitride hollow tubes exhibited superior photodegradation activities for methylene blue and phenol thanks to the content of nitrogen impurities nitrogen, a unique hollow structure, and a larger specific surface area; porosity directly correlated with reaction kinetics [43]. The combination of advanced oxidation processes and photocatalysis can further enhance the photodegradation activity of g-C3N4. An et al. reported a single-atom Fe g-C3N4 catalyst with Fe (II)-Nx active sites to accelerate the production of HO· radicals, leading to excellent degradation efficiency for various organics (MB, MO, RhB, and phenol) [100].

3.1.2. Heavy Metal Detoxification

Toxic heavy metal ions in industrial wastewater can also be removed by g-C3N4’s dual adsorption-photoreduction capability [102,103,104,105]. In order to enhance the charge migration, Lei et al. coupled carbon nitride nanosheets with MIL-88B(Fe), which possessedthe highest catalytic activity to reduce Cr (VI) under visible light [105]. Three-dimensional g-C3N4@cellulose aerogel is certified to remove hexavalent chromium and antibiotics simultaneously under light irradiation, where hierarchical pores enable simultaneous adsorption and reduction [106]. Pd nanocones supported on g-C3N4 showed enhanced catalytic reduction of hexavalent chromium under visible-light irradiation [107].

3.1.3. Air Purification

The removal rate of NO commonly used to investigate the degradation capacity of C3N4 in air [108,109,110,111,112]. Li et al. prepared an illite particle-modified g-C3N4 to enhance the photocatalytic NO removal activity of g-C3N4 [113]. The formation of heterojunctions between illite and g-C3N4 accelerated the charge migration, leading toexcellent photocatalytic NO removal activity. Porous g-C3N4 microtubes with N-vacancies showed 2.6 times higher NO removal than bulk material, and the 3D porous wall structure facilitated gas diffusion and exposure of active sites [39]. The existence of N-vacancies can improve the adsorption ability for NO, the charge capture, and the light-absorbing capability of g-C3N4. Meanwhile, the diffusion of reactants and the transfer of charge are promoted by the porous wall structure. In addition, the preparation of 3D materials is considered an efficient approach for air purification. Hu et al. modified g-C3N4 with perylene imide (PI) and graphene oxide (GO) to prepare a g-C3N4-based aerogel [112]. The excellent activity in NO removal that was observed was attributed to the strong light absorption and favorable charge transport.

3.1.4. Outlook

In the environmental research, the application of g-C3N4 materials in degrading organic pollutants in water is still a popular research direction. It is worth noting that the targets of degradation have shifted from organic dyes to emerging refractory pollutants (antibiotics, phenolic organic matter) (Figure 12a). Nonetheless, there are still many challenges relating to the practical application of g-C3N4 materials for the degradation of pollutants in water, such as the pathway and mechanism of degradation, limited real wastewater validation, and the trade-off between SSA enhancement and quantum efficiency loss. Future work must resolve this fundamental trade-off; SSA expansion often introduces defect-induced recombination centers, demanding balanced design to preserve quantum efficiency.

3.2. Energy Conversion and Storage: Bridging Photocatalysis to Practical Energy Systems

It is well known that the photo-generated electrons of g-C3N4 have a certain reduction ability due to the location of the appropriate conduction band, which can be adjusted by various approaches. Since they were found to be able to produce hydrogen, g-C3N4 materials have been widely used in many fields of energy regeneration, including hydrogen production, CO2 reduction, nitrogen fixation, and so on (Table 9) [25,26,28,36,38,46,51,53,54,55,58,60,65,69,71,74,77,81,82,84,114,115,116,117,118,119]. Enhanced SSA improves energy conversion efficiency by exposing more catalytic sites, shortening charge migration paths, and facilitating mass transport.

3.2.1. Solar Hydrogen Production

As the most ideal renewable energy, hydrogen energy has the advantages of clean, efficient, and convenient storage. Photocatalytic water splitting technology is a commonly used green method for hydrogen production. The reaction process is shown in the following equation:
4 H + + 4 e 2 H 2
H 2 O + 2 h + O 2 + 2 H + + 4 e
The precondition of photocatalytic reaction is that the potential of the valence band in the catalyst is more positive than that of O2/H2O (1.23 V vs. NHE), and its potential of conduction band is more negative than that of H+/H2 (0 V vs. NHE). In the past few years, g-C3N4 has been widely used in photocatalytic hydrogen production, due to its suitable conduction band [28,120,121,122]. The purpose of modification is to increase the number of photogenerated electrons in the system to support the reduction reaction. Hollow g-C3N4 nanotubes with an eight-fold increase in specific surface area prepared by precursor mixing were applied to product hydrogen with Pt as the cocatalyst, and exhibited a higher photocatalytic production rate of 1073.6 µmol·h−1·g−1 [28]. The enhanced photocatalytic H2 production benefited from the porous and nanotubular structure, which facilitated charge carrier migration and separation. Furthermore, Che et al. enhanced the overall water splitting of g-C3N4 by preparing an in-plane (C ring)-C3N4 heterostructure with fast spatial transfer of photo-generated electrons [122]. Single-atom loading is an efficient approach to further enhance the hydrogen production rate of g-C3N4. For example, single-atom Cu decorated tubular g-C3N4 exhibited a superior visible-light photocatalytic hydrogen production rate (≈212 µmol h−1/0.02 g catalyst), which was ascribed to the improved in-plane and interlayer separation/transfer of the charge carriers [123].

3.2.2. CO2 Photoreduction

Recently, photocatalytic technology has been applied in CO2 photoreduction to produce hydrocarbon fuels. This technology has become a research hotspot due to its bright prospects for alleviating the greenhouse effect and energy crisis simultaneously. The basic steps include the adsorption of CO2, the production and migration of carriers, the reduction reaction, and the desorption of products. The main evaluation parameters of CO2 photoreduction performance are conversion rate and selectivity.
As a hot semiconductor, g-C3N4 has been considered a potential catalyst for CO2 reduction [124,125,126,127]. Nevertheless, the activity of bulk g-C3N4 in CO2 reduction remains inhibited by their small specific surface area, high recombination of photo-generated carriers, and low electron conductivity. A 3D porous g-C3N4/C nanosheets composite was successfully prepared to achieve the excellent CO2 reduction activity [128]. Owing to its enhanced light trapping/utilization, highly efficient CO2 adsorption ability by mesopores, and low recombination of photogenerated carriers, it showed better CO2 reduction activity than that of bulk g-C3N4 (CO and CH4 yield of 229 and 112 μmol g−1-cat). Moreover, single Ni atoms have been anchored on porous few-layer g-C3N4 to improve photocatalytic CO2 reduction [129]. Thanks to the synergistic N-Ni-N connection and interfacial carrier transfer, Ni-C3N4 exhibited a CO generation rate of 8.6 µmol g−1 h−1 under visible light illumination, which was 7.8 times that of the porous few-layer g-C3N4 (1.1 µmol g−1 h−1). Samanta et al. enhanced the photocatalytic CO2 reduction of g-C3N4 by preparing doped g-C3N4 [124]. The C, O co-doped g-C3N4 exhibited a higher yield of CH3OH (4.18 mmol g−1 in 6 h) than bulk C3N4 (2.8 mmol g−1).

3.2.3. Nitrogen Fixation

Nitrogen fixation, a potential pathway for ammonia synthesis, has attracted wide attention in the field of energy as a solution for the energy and environmental problems caused by traditional ammonia synthesis. The whole process is a multi-step reaction between photo-generated electrons and protons.
H 2 O 1 2 O 2 + 2 H + + 2 e
2 H + + 2 e H 2
N 2 + e 2 N 2
N 2 + H + + e N 2 H
N 2 + 2 H + + 2 e N 2 H 2
N 2 + 4 H + + 4 e N 2 H 4
N 2 + 5 H + + 4 e N 2 H 5 +
N 2 + 6 H + + 6 e 2 N H 3
N 2 + 8 H + + 6 e 2 N H 4 +
It can be seen that this process requires a large number of electrons, competes with hydrogen production, and involves intermediates and by-products. Despite thermodynamic challenges (N≡N bond dissociation: 941 kJ mol−1), some advances have been made [130,131,132]. Qiu et al. combined black phosphorus nanosheets with the g-C3N4 nanosheets to enhance performance in visible-light nitrogen photo-fixation [132]. The formation of C-P covalent bonds increased the number of excited electrons and facilitated the separation efficiency of carriers. Additionally, the obtained ultrathin sulfur-doped g-C3N4 porous nanosheets with large lateral size and carbon vacancies had a photocatalytic nitrogen fixation rate of 5.99 mMh−1 gCat−1 (2.8 times higher than that of bulk g-C3N4) due to the carbon vacancies and edge sites synergizing for N₂ activation [133]. Yin et al. synthesized single-atom electrocatalysts consisting of transition metal atoms and monolayer g-C3N4 for NH3 production [134]. For single-atom Pt-C3N4, the adsorption capacity of N2 is higher than that of H atoms, suggesting excellent nitrogen-fixation activity.

3.2.4. Energy Storage

In addition to the above, g-C3N4-based composites have been applied in energy storage [135,136,137,138]. For instance, Wang et al. introduced g-C3N4 into a three-dimensional hierarchical porous graphene to achieve excellent cycling performance for sulfur cathodes in lithium–sulfur batteries [137]. In lithium-ion batteries, the prepared layered g-C3N4@reduced graphene oxide composites possessed excellent cycle stability (899.3 mAh g−1 after 350 cycles under 500 mA g−1) and remarkable rate performance (595.1 mAh g−1 after 1000 cycles under 1000 mA g−1), because of the large interlayer distances, rich N-active sites, and a microporous structure accommodating volume expansion [138]. Zhang et al. prepared a 3D porous sulfur/graphene@g-C3N4 hybrid sponge and used it as a cathode for Li-S batteries [135]. The superior high-rate capability (612 mAh g−1 at 10 C) benefited from the numerous adhesion sites of polysulfides and the efficient electron/Li+ transport pathways.

3.2.5. Outlook

As shown in Figure 12b, the research on CO2 reduction and nitrogen fixation in addition to hydrogen production is increasing gradually. The further development of these directions is of great significance to alleviate the energy crisis. In order to prepare g-C3N4 materials with excellent performance, further exploration of reaction mechanisms and pathways is necessary. While CO2RR and N2 fixation show promise, scalability requires breakthroughs in catalyst stability under continuous operation and product-separation technologies for mixed-output systems. Scalability requires stabilizing high-SSA architectures against restructuring during continuous operation.

3.3. Biomedical Innovations: Antimicrobial and Sensing Platforms

3.3.1. Photodynamic Antimicrobial Therapy

There are many microorganisms which may be harmful to the human body. The antibacterial process uses the active substances produced by photocatalyst, such as superoxide anion and hydrogen peroxide, to inhibit the reproduction of microorganisms and hinder their internal reactions. Therefore, the antibacterial properties of g-C3N4 materials depend on its ability to produce active substances [139,140,141,142]. g-C3N4 nanosheets have attached increasing attention for use in photocatalytic disinfection, due to their enhanced optical, mechanical, and electrical properties. Li et al. assembled functionalized g-C3N4 nanosheets composite membranes with superior self-cleaning and antimicrobial properties [140]. Kang et al. developed high-quality g-C3N4 nanosheets with 82.61 m2/g via a bacterial etching approach, obtaining better photocatalytic disinfection performance compared with bulk g-C3N4 [139]. In addition, a g-C3N4/perylene-3,4,9,10-tetracarboxylic diimide (PDINH) heterostructure was synthesized to increase the number of reactive oxygen species, leading to an excellent inactivation effect on bacteria [141].

3.3.2. Biosensing and Diagnostics

Another application of g-C3N4 materials in the field of biology is photo-electrochemical sensing, which depends on visible-light response and the efficiency of charge carrier separation [63,143,144,145,146]. For example, the core–shell LaFeO3@g-C3N4 p-n heterostructure showed improved photoelectrochemical performance for streptomycin sensing, due to the wider absorption band edge and stronger photocurrent signal [146]. Ultrathin g-C3N4 nanosheets have been used as heparin sensing platform with a heparin detection limit of 18 ng mL−1 [63]. In addition, Chen et al. tested DNA methyltransferase activity using a g-C3N4 nanosheet electrochemiluminescence biosensor, which exhibited an ultralow detection limit down to 0.043 U mL−1 [143].

3.3.3. Outlook

It can be seen that the wide application of g-C3N4 materials in the biological field can bring great benefits to humans in terms of medical and health care. However, several challenges persist. These include concerns over biocompatibility in in vivo applications, the paucity of research on long-term cytotoxicity, and the difficulties associated with the scalable synthesis of medical-grade g-C3N4. Increased SSA enhances biomedical functionality by amplifying reactive oxygen species (ROS) generation per unit mass and improving biomolecule adhesion.

4. Conclusions and Outlook

The progress in research into g-C3N4 with large specific surface areas has been summarized and discussed in this work, involving the applicationand controllable synthesis of high-SSA g-C3N4. Several applications of g-C3N4 materials have been presented. It can be seen that the research into g-C3N4 has involved many fields relating to energy and the environment, including hydrogen production, degradation, energy storage, sensing, and so on. Nevertheless, the application of g-C3N4 in some fields has not been deeply explored. Thus, g-C3N4 is still a potential material to mitigate energy and environmental crises. Although the performance of g-C3N4 has significantly improved in recent years, there are still challenges to be addressed in the exploration of g-C3N4.
(i)
For the design and synthesis of g-C3N4 materials with excellent performance, overall consideration should be given to the specific surface area, band structure, and defects of g-C3N4 materials. For photocatalysis, the realization of excellent photocatalytic efficiency requires the combined efforts of the light absorption ability, charge transfer, and separation ability, and adsorption ability for the catalyst’s reactants. Mechanisms to promote the properties of g-C3N4-derived materials should be explored. The further functionalization mentioned in this work can improve the performances of g-C3N4 by promoting the transfer and separation of charge. However, the exact transfer pathway of charge in g-C3N4 derived material is still ambiguous or unsubstantiated. Further exploration of the mechanism is of great significance for the synthesis of g-C3N4 materials with target structures. Combination of improved methods is required. The published methods for modification of g-C3N4 have usually aimed to improve a particular aspect of performance. Therefore, the synthesis of high-performance g-C3N4 can be realized by combining various modification methods according to the requirements of specific application fields. The design and assembly of material reactors should be considered. Generally, most prepared g-C3N4 materials are powder, which is not suitable for all applications. The design of suitable reactors can meet the particular requirements of some reactions to broaden the application range of g-C3N4.
(ii)
Further exploration of reaction mechanisms should include interface reactions between reactants and g-C3N4 materials. Many applied reactions of g-C3N4 materials have occurred in water or other media where the reaction isdirectly affected by the interface effect between reactants and materials. Regarding the pathways of g-C3N4 material reactions with reactants, our experimental process was a macroscopic process, so that the specific reaction path between g-C3N4 materials and the reactants could not be fully assessed. Such exploration is very important for the targeted synthesis and application of high-performance g-C3N4 materials. Theoretical calculations related to g-C3N4 materials can be combined with experimental results to clarify the specific active mechanisms of g-C3N4 materials, which is conducive to understanding the properties of materials and the occurrence of reactions. However, the direction of calculation needs to be further explored.
(iii)
Application fields of g-C3N4 materials include practical industrial application of g-C3N4 materials in “old” fields. The degradation activity of the g-C3N4 material under actual water conditions should be studied, including the coexistence of various pollutants, fluid medium, sunlight irradiation, and so on. With regard to improving the performances of g-C3N4 material in “new” fields, g-C3N4 materials have been subject to a relatively few studies in the fields of nitrogen fixation, antibacterial activity, and sensing, leaving large room for improvement.
Thus, g-C3N4 materials with target performances and structures can be designed based on the reaction mechanism to realize their value in the fields of energy and the environment in the future.

Funding

This work was funded by the Research Program of Qilu Institute of Technology (grant number QIT24TP001 and QIT24TP004).

Acknowledgments

This work was supported by the Research Program of Qilu Institute of Technology (QIT24TP001 and QIT24TP004) and the Open Project of the Key Laboratory of Ultralight Materials and Surface Technology of the Ministry of Education.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tong, T.; Zhang, M.; Chen, W.; Huo, X.; Xu, F.; Yan, H.; Lai, C.; Wang, W.; Hu, S.; Qin, L.; et al. Recent advances in carbon-based material semiconductor composite photoelectrocatalysts: Synthesis, improvement strategy, and organic pollutant removal. Coord. Chem. Rev. 2024, 500, 215498. [Google Scholar] [CrossRef]
  2. Tang, L.; Zou, J. p-type two-dimensional semiconductors: From materials preparation to electronic applications. Nano-Micro Lett. 2023, 15, 230. [Google Scholar] [CrossRef]
  3. Liu, L.; Bai, B.; Yang, X.; Du, Z.; Jia, G. Anisotropic heavy-metal-free semiconductor nanocrystals: Synthesis, properties, and applications. Chem. Rev. 2023, 123, 3625–3692. [Google Scholar] [CrossRef]
  4. Liras, M.; Barawi, M.; de la Peña O’Shea, V.A. Hybrid materials based on conjugated polymers and inorganic semiconductors as photocatalysts: From environmental to energy applications. Chem. Soc. Rev. 2019, 48, 5454–5487. [Google Scholar] [CrossRef]
  5. Zhou, X.; Wang, S.; Li, Y.; Yang, Y.; Xiao, X.; Chen, G. Boosting Li-metal anode performance with lithiophilic Li-Zn seeds in a 2D reduced graphene oxide scaffold. Adv. Energy Mater. 2024, 15, 2403640. [Google Scholar] [CrossRef]
  6. Ning, M.; Wang, Y.; Wu, L.; Yang, L.; Chen, Z.; Song, S.; Yao, Y.; Bao, J.; Chen, S.; Ren, Z. Hierarchical interconnected NiMoN with large specific surface area and high mechanical strength for efficient and stable alkaline water/seawater hydrogen evolution. Nano-Micro Lett. 2023, 15, 157. [Google Scholar] [CrossRef] [PubMed]
  7. Kuang, P.; Sayed, M.; Fan, J.; Cheng, B.; Yu, J. 3D Graphene-based H2-production photocatalyst and electrocatalyst. Adv. Energy Mater. 2020, 10, 1903802. [Google Scholar] [CrossRef]
  8. Zang, Y.; Pei, F.; Huang, J.; Fu, Z.; Xu, G.; Fang, X. Large-area preparation of crack-free crystalline microporous conductive membrane to upgrade high energy lithium-sulfur batteries. Adv. Energy Mater. 2018, 8, 1802052. [Google Scholar] [CrossRef]
  9. Lu, S.Y.; Jin, M.; Zhang, Y.; Niu, Y.B.; Gao, J.C.; Li, C.M. Chemically exfoliating biomass into a graphene-like porous active carbon with rational pore structure, good conductivity, and large surface area for high-performance supercapacitors. Adv. Energy Mater. 2017, 8, 1702545. [Google Scholar] [CrossRef]
  10. 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]
  11. Zhang, X.; Zhu, K.; Xie, C.; Yang, P. Vertically implanting MoSe2 nanosheets on superior thin C-doped g-C3N4 nanosheets towards interface-enhanced electrochemical activities. Carbon 2024, 220, 118884. [Google Scholar] [CrossRef]
  12. Yang, H.; Zhang, A.; Ding, J.; Hu, R.; Gong, Y.; Li, X.; Chen, L.; Chen, P.; Tian, X. Amino modulation on the surface of graphitic carbon nitride for enhanced photocatalytic hydrogen production. Carbon 2024, 219, 118841. [Google Scholar] [CrossRef]
  13. Xu, T.; Hur, J.; Niu, P.; Wang, S.; Lee, S.; Chun, S.-E.; Li, L. Synthesis of crystalline g-C3N4 with rock/molten salts for efficient photocatalysis and piezocatalysis. Green Energy Environ. 2024, 9, 890–898. [Google Scholar] [CrossRef]
  14. Ding, X.; Gao, R.; Chen, Y.; Wang, H.; Liu, Y.; Zhou, B.; Wang, C.; Bai, G.; Qiu, W. Carbon vacancies in graphitic carbon nitride-driven high catalytic performance of Pd/CN for phenol-selective hydrogenation to cyclohexanone. ACS Catal. 2024, 14, 3308–3319. [Google Scholar] [CrossRef]
  15. Wang, W.; He, Q.; Yi, Y.; Xiao, Y.; Xiao, X.; Yang, H.; Dong, X. Boosting piezocatalytic activity of graphitic carbon nitride for degrading antibiotics through morphologic regulation and chlorine doping. J. Clean. Prod. 2023, 415, 137818. [Google Scholar] [CrossRef]
  16. Tabarkhoon, F.; Abolghasemi, H.; Rashidi, A.; Bazmi, M.; Alivand, M.S.; Tabarkhoon, F.; Farahani, M.V.; Esrafili, M.D. Synthesis of novel and tunable micro-mesoporous carbon nitrides for ultra-high CO2 and H2S capture. Chem. Eng. J. 2023, 456, 140973. [Google Scholar] [CrossRef]
  17. 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]
  18. Wang, L.; Wang, C.; Hu, X.; Xue, H.; Pang, H. Metal/graphitic carbon nitride composites: Synthesis, structures, and applications. Chem.-Asian J. 2016, 11, 3305–3328. [Google Scholar] [CrossRef]
  19. Tang, J.; Xu, R.; Sui, G.; Guo, D.; Zhao, Z.; Fu, S.; Yang, X.; Li, Y.; Li, J. Double-shelled porous g-C3N4 nanotubes modified with amorphous Cu-doped FeOOH nanoclusters as 0D/3D non-homogeneous photo-Fenton catalysts for effective removal of organic dyes. Small 2023, 19, 2208232. [Google Scholar] [CrossRef]
  20. Song, Y.; Zhou, C.; Zheng, Z.; Sun, P.; She, Y.; Huang, F.; Mo, Z.; Yuan, J.; Li, H.; Xu, H. Porous carbon nitride nanotubes efficiently promote two-electron O2 reduction for photocatalytic H2O2 production. J. Alloys Compd. 2023, 934, 167901. [Google Scholar] [CrossRef]
  21. Liu, D.; Li, C.; Zhao, C.; Zhao, Q.; Niu, T.; Pan, L.; Xu, P.; Zhang, F.; Wu, W.; Ni, T. Facile synthesis of three-dimensional hollow porous carbon doped polymeric carbon nitride with highly efficient photocatalytic performance. Chem. Eng. J. 2022, 438, 135623. [Google Scholar] [CrossRef]
  22. Li, Q.; Tong, Y.; Zeng, Y.; Gu, X.-K.; Ding, M. Carbohydrate-regulated synthesis of ultrathin porous nitrogen-vacancy polymeric carbon nitride for highly efficient visible-light hydrogen evolution. Chem. Eng. J. 2022, 450, 138010. [Google Scholar] [CrossRef]
  23. Attri, P.; Garg, P.; Sharma, P.; Singh, R.; Chauhan, M.; Lim, D.-K.; Kumar, S.; Chaudhary, G.R. Precursor-dependent fabrication of exfoliated graphitic carbon nitride (gCN) for enhanced photocatalytic and antimicrobial activity under visible light irradiation. J. Clean. Prod. 2023, 422, 138538. [Google Scholar] [CrossRef]
  24. Zhang, J.; Chen, Y.; Wang, X. Two-dimensional covalent carbon nitride nanosheets: Synthesis, functionalization, and applications. Energy Environ. Sci. 2015, 8, 3092–3108. [Google Scholar] [CrossRef]
  25. Wen, J.; Zhang, S.; Liu, Y.; Zhai, Y. Formic acid assisted fabrication of Oxygen-doped Rod-like carbon nitride with improved photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2022, 624, 338–347. [Google Scholar] [CrossRef]
  26. Wang, H.; Jiang, J.; Yu, L.; Peng, J.; Song, Z.; Xiong, Z.; Li, N.; Xiang, K.; Zou, J.; Hsu, J.P.; et al. Tailoring advanced N-defective and S-doped g-C3N4 for photocatalytic H2 evolution. Small 2023, 19, 2301116. [Google Scholar] [CrossRef] [PubMed]
  27. Li, H.; Qiao, S.; Zhang, H.; Qiao, Y.; Liu, J.; Li, Y. Highly sensitive and selective demethylase FTO detection using a DNAzyme-mediated CRISPR/Cas12a signal cascade amplification electrochemiluminescence biosensor with C-CN/PCNV heterojunction as emitter. Biosens. Bioelectron. 2024, 256, 116276. [Google Scholar] [CrossRef]
  28. Wang, X.; Zhou, C.; Shi, R.; Liu, Q.; Waterhouse, G.I.N.; Wu, L.; Tung, C.-H.; Zhang, T. Supramolecular precursor strategy for the synthesis of holey graphitic carbon nitride nanotubes with enhanced photocatalytic hydrogen evolution performance. Nano Res. 2019, 12, 2385–2389. [Google Scholar] [CrossRef]
  29. Chen, X.; Shi, R.; Chen, Q.; Zhang, Z.; Jiang, W.; Zhu, Y.; Zhang, T. Three-dimensional porous g-C3N4 for highly efficient photocatalytic overall water splitting. Nano Energy 2019, 59, 644–650. [Google Scholar] [CrossRef]
  30. Barrio, J.; Shalom, M. Rational design of carbon nitride materials by supramolecular preorganization of monomers. ChemCatChem 2018, 10, 5573–5586. [Google Scholar] [CrossRef]
  31. Bao, J.; Bai, W.; Wu, M.; Gong, W.; Yu, Y.; Zheng, K.; Liu, L. Template-mediated copper doped porous g-C3N4 for efficient photodegradation of antibiotic contaminants. Chemosphere 2022, 293, 133607. [Google Scholar] [CrossRef] [PubMed]
  32. Dong, G.; Ho, W.; Li, Y.; Zhang, L. Facile synthesis of porous graphene-like carbon nitride (C6N9H3) with excellent photocatalytic activity for NO removal. Appl. Catal. B Environ. 2015, 174–175, 477–485. [Google Scholar] [CrossRef]
  33. Zhang, X.-S.; Hu, J.-Y.; Jiang, H. Facile modification of a graphitic carbon nitride catalyst to improve its photoreactivity under visible light irradiation. Chem. Eng. J. 2014, 256, 230–237. [Google Scholar] [CrossRef]
  34. Zhong, Y.; Wang, Z.; Feng, J.; Yan, S.; Zhang, H.; Li, Z.; Zou, Z. Improvement in photocatalytic H2 evolution over g-C3N4 prepared from protonated melamine. Appl. Surf. Sci. 2014, 295, 253–259. [Google Scholar] [CrossRef]
  35. Yan, H.; Chen, Y.; Xu, S. Synthesis of graphitic carbon nitride by directly heating sulfuric acid treated melamine for enhanced photocatalytic H2 production from water under visible light. Int. J. Hydrogen Energy 2012, 37, 125–133. [Google Scholar] [CrossRef]
  36. Han, Q.; Cheng, Z.; Wang, B.; Zhang, H.; Qu, L. Significant enhancement of visible-light-driven hydrogen evolution by structure regulation of carbon nitrides. ACS Nano 2018, 12, 5221–5227. [Google Scholar] [CrossRef]
  37. Wu, S.; Wen, S.; Xu, X.; Huang, G.; Cui, Y.; Li, J.; Qu, A. Facile synthesis of porous graphene-like carbon nitride nanosheets with high surface area and enhanced photocatalytic activity via one-step catalyst-free solution self-polymerization. Appl. Surf. Sci. 2018, 436, 424–432. [Google Scholar] [CrossRef]
  38. Mo, Z.; Xu, H.; Chen, Z.; She, X.; Song, Y.; Wu, J.; Yan, P.; Xu, L.; Lei, Y.; Yuan, S.; et al. Self-assembled synthesis of defect-engineered graphitic carbon nitride nanotubes for efficient conversion of solar energy. Appl. Catal. B Environ. 2018, 225, 154–161. [Google Scholar] [CrossRef]
  39. Wang, Z.; Huang, Y.; Chen, M.; Shi, X.; Zhang, Y.; Cao, J.; Ho, W.; Lee, S.C. Roles of N-vacancies over porous g-C3N4 microtubes during photocatalytic NOx removal. ACS Appl. Mater. Interfaces 2019, 11, 10651–10662. [Google Scholar] [CrossRef]
  40. Cheng, J.; Hu, Z.; Li, Q.; Li, X.; Fang, S.; Wu, X.; Li, M.; Ding, Y.; Liu, B.; Yang, C.; et al. Fabrication of high photoreactive carbon nitride nanosheets by polymerization of amidinourea for hydrogen production. Appl. Catal. B Environ. 2019, 245, 197–206. [Google Scholar] [CrossRef]
  41. Mondal, S.; Mark, G.; Tashakory, A.; Volokh, M.; Shalom, M. Porous carbon nitride rods as an efficient photoanode for water splitting and benzylamine oxidation. J. Mater. Chem. A 2024, 12, 11502–11510. [Google Scholar] [CrossRef]
  42. Zhang, D.; Guo, Y.; Zhao, Z. Porous defect-modified graphitic carbon nitride via a facile one-step approach with significantly enhanced photocatalytic hydrogen evolution under visible light irradiation. Appl. Catal. B Environ. 2018, 226, 1–9. [Google Scholar] [CrossRef]
  43. Tian, L.; Li, J.; Liang, F.; Wang, J.; Li, S.; Zhang, H.; Zhang, S. Molten salt synthesis of tetragonal carbon nitride hollow tubes and their application for removal of pollutants from wastewater. Appl. Catal. B Environ. 2018, 225, 307–313. [Google Scholar] [CrossRef]
  44. Vinu, A.; Ariga, K.; Mori, T.; Nakanishi, T.; Hishita, S.; Golberg, D.; Bando, Y. Preparation and characterization of well-ordered hexagonal mesoporous carbon nitride. Adv. Mater. 2005, 17, 1648–1652. [Google Scholar] [CrossRef]
  45. Vinu, A. Two-dimensional hexagonally-ordered mesoporous carbon nitrides with tunable pore diameter, surface area and nitrogen content. Adv. Funct. Mater. 2008, 18, 816–827. [Google Scholar] [CrossRef]
  46. Chen, X.; Jun, Y.-S.; Takanabe, K.; Maeda, K.; Domen, K.; Fu, X.; Antonietti, M.; Wang, X. Ordered mesoporous SBA-15 type graphitic carbon nitride: A semiconductor host structure for photocatalytic hydrogen evolution with visible light. Chem. Mater. 2009, 21, 4093–4095. [Google Scholar] [CrossRef]
  47. Cui, Y.; Zhang, J.; Zhang, G.; Huang, J.; Liu, P.; Antonietti, M.; Wang, X. Synthesis of bulk and nanoporous carbon nitride polymers from ammonium thiocyanate for photocatalytic hydrogen evolution. J. Mater. Chem. 2011, 21, 13032. [Google Scholar] [CrossRef]
  48. Zhao, Z.; Dai, Y.; Lin, J.; Wang, G. Highly-ordered mesoporous carbon nitride with ultrahigh surface area and pore volume as a superior dehydrogenation catalyst. Chem. Mater. 2014, 26, 3151–3161. [Google Scholar] [CrossRef]
  49. Wang, F.; Feng, Y.; Chen, P.; Wang, Y.; Su, Y.; Zhang, Q.; Zeng, Y.; Xie, Z.; Liu, H.; Liu, Y.; et al. Photocatalytic degradation of fluoroquinolone antibiotics using ordered mesoporous g-C3N4 under simulated sunlight irradiation: Kinetics, mechanism, and antibacterial activity elimination. Appl. Catal. B Environ. 2018, 227, 114–122. [Google Scholar] [CrossRef]
  50. Srinivasu, P.; Vinu, A.; Hishita, S.; Sasaki, T.; Ariga, K.; Mori, T. Preparation and characterization of novel microporous carbon nitride with very high surface area via nanocasting technique. Microporous Mesoporous Mater. 2008, 108, 340–344. [Google Scholar] [CrossRef]
  51. Kailasam, K.; Epping, J.D.; Thomas, A.; Losse, S.; Junge, H. Mesoporous carbon nitride-silica composites by a combined sol-gel/thermal condensation approach and their application as photocatalysts. Energy Environ. Sci. 2011, 4, 4668. [Google Scholar] [CrossRef]
  52. Shiraishi, Y.; Kofuji, Y.; Sakamoto, H.; Tanaka, S.; Ichikawa, S.; Hirai, T. Effects of surface defects on photocatalytic H2O2 production by mesoporous graphitic carbon nitride under visible light irradiation. ACS Catal. 2015, 5, 3058–3066. [Google Scholar] [CrossRef]
  53. Tong, Z.; Yang, D.; Li, Z.; Nan, Y.; Ding, F.; Shen, Y.; Jiang, Z. Thylakoid-inspired multishell g-C3N4 nanocapsules with enhanced visible-light harvesting and electron transfer properties for high-efficiency photocatalysis. ACS Nano 2017, 11, 1103–1112. [Google Scholar] [CrossRef]
  54. Lin, W.; Hong, W.; Sun, L.; Yu, D.; Yu, D.; Chen, X. Bioinspired mesoporous chiral nematic graphitic carbon nitride photocatalysts modulated by polarized light. ChemSusChem 2018, 11, 114–119. [Google Scholar] [CrossRef]
  55. Liu, X.; Pang, F.; He, M.; Ge, J. Confined reaction inside nanotubes: New approach to mesoporous g-C3N4 photocatalysts. Nano Res. 2017, 10, 3638–3647. [Google Scholar] [CrossRef]
  56. Cui, L.; Song, J.; McGuire, A.F.; Kang, S.; Fang, X.; Wang, J.; Yin, C.; Li, X.; Wang, Y.; Cui, B. Constructing highly uniform onion-ring-like graphitic carbon nitride for efficient visible-light-driven photocatalytic hydrogen evolution. ACS Nano 2018, 12, 5551–5558. [Google Scholar] [CrossRef]
  57. Si, Y.; Huang, T.; Li, Q.; Huang, Y.; Gao, S.P.; Chen, M.; Wu, L. Hierarchical macro-mesoporous polymeric carbon nitride microspheres with narrow bandgap for enhanced photocatalytic hydrogen production. Adv. Mater. Interfaces 2018, 5, 1801241. [Google Scholar] [CrossRef]
  58. Zhang, J.; Zhang, M.; Yang, C.; Wang, X. Nanospherical carbon nitride frameworks with sharp edges accelerating charge collection and separation at a soft photocatalytic interface. Adv. Mater. 2014, 26, 4121–4126. [Google Scholar] [CrossRef] [PubMed]
  59. Fan, B.; Xing, L.; He, Q.; Zhou, F.; Yang, X.; Wu, T.; Tong, G.; Wang, D.; Wu, W. Selective synthesis and defects steering superior microwave absorption capabilities of hollow graphitic carbon nitride micro-polyhedrons. Chem. Eng. J. 2022, 435, 135086. [Google Scholar] [CrossRef]
  60. Shi, L.; Chang, K.; Zhang, H.; Hai, X.; Yang, L.; Wang, T.; Ye, J. Drastic enhancement of photocatalytic activities over phosphoric acid protonated porous g-C3N4 nanosheets under visible light. Small 2016, 12, 4431–4439. [Google Scholar] [CrossRef]
  61. Du, X.; Zou, G.; Wang, Z.; Wang, X. A scalable chemical route to soluble acidified graphitic carbon nitride: An ideal precursor for isolated ultrathin g-C3N4 nanosheets. Nanoscale 2015, 7, 8701–8706. [Google Scholar] [CrossRef] [PubMed]
  62. Qian, J.; Yuan, A.; Yao, C.; Liu, J.; Li, B.; Xi, F.; Dong, X. Highly efficient photo-reduction of p-Nitrophenol by protonated graphitic carbon nitride nanosheets. ChemCatChem 2018, 10, 4747–4754. [Google Scholar] [CrossRef]
  63. Ma, T.Y.; Tang, Y.; Dai, S.; Qiao, S.Z. Proton-functionalized two-dimensional graphitic carbon nitride nanosheet: An excellent metal-/label-free biosensing platform. Small 2014, 10, 2382–2389. [Google Scholar] [CrossRef]
  64. Yuan, X.; Zhou, C.; Jin, Y.; Jing, Q.; Yang, Y.; Shen, X.; Tang, Q.; Mu, Y.; Du, A.K. Facile synthesis of 3D porous thermally exfoliated g-C3N4 nanosheet with enhanced photocatalytic degradation of organic dye. J. Colloid Interface Sci. 2016, 468, 211–219. [Google Scholar] [CrossRef]
  65. 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. [Google Scholar] [CrossRef]
  66. Teng, Z.; Lv, H.; Wang, C.; Xue, H.; Pang, H.; Wang, G. Bandgap engineering of ultrathin graphene-like carbon nitride nanosheets with controllable oxygenous functionalization. Carbon 2017, 113, 63–75. [Google Scholar] [CrossRef]
  67. Feng, J.; Chen, T.; Liu, S.; Zhou, Q.; Ren, Y.; Lv, Y.; Fan, Z. Improvement of g-C3N4 photocatalytic properties using the Hummers method. J. Colloid Interface Sci. 2016, 479, 1–6. [Google Scholar] [CrossRef]
  68. Zhang, X.; Xie, X.; Wang, H.; Zhang, J.; Pan, B.; Xie, Y. Enhanced photoresponsive ultrathin graphitic-phase C3N4 nanosheets for bioimaging. J. Am. Chem. Soc. 2013, 135, 18–21. [Google Scholar] [CrossRef]
  69. Niu, P.; Zhang, L.; Liu, G.; Cheng, H.-M. Graphene-like carbon nitride nanosheets for improved photocatalytic activities. Adv. Funct. Mater. 2012, 22, 4763–4770. [Google Scholar] [CrossRef]
  70. Luo, B.; Song, R.; Jing, D. Significantly enhanced photocatalytic hydrogen generation over graphitic carbon nitride with carefully modified intralayer structures. Chem. Eng. J. 2018, 332, 499–507. [Google Scholar] [CrossRef]
  71. Yang, S.; Gong, Y.; Zhang, J.; Zhan, L.; Ma, L.; Fang, Z.; Vajtai, R.; Wang, X.; Ajayan, P.M. Exfoliated graphitic carbon nitride nanosheets as efficient catalysts for hydrogen evolution under visible light. Adv. Mater. 2013, 25, 2452–2456. [Google Scholar] [CrossRef]
  72. Lin, Q.; Li, L.; Liang, S.; Liu, M.; Bi, J.; Wu, L. Efficient synthesis of monolayer carbon nitride 2D nanosheet with tunable concentration and enhanced visible-light photocatalytic activities. Appl. Catal. B Environ. 2015, 163, 135–142. [Google Scholar] [CrossRef]
  73. Gao, M.; Feng, J.; Zhang, Z.; Gu, M.; Wang, J.; Zeng, W.; Lv, Y.; Ren, Y.; Wei, T.; Fan, Z. Wrinkled ultrathin graphitic C3N4 nanosheets for photocatalytic degradation of organic wastewater. ACS Appl. Nano Mater. 2018, 1, 6733–6741. [Google Scholar] [CrossRef]
  74. Zhang, Z.; Zhang, Y.; Lu, L.; Si, Y.; Zhang, S.; Chen, Y.; Dai, K.; Duan, P.; Duan, L.; Liu, J. Graphitic carbon nitride nanosheet for photocatalytic hydrogen production: The impact of morphology and element composition. Appl. Surf. Sci. 2017, 391, 369–375. [Google Scholar] [CrossRef]
  75. Nie, H.; Ou, M.; Zhong, Q.; Zhang, S.; Yu, L. Efficient visible-light photocatalytic oxidation of gaseous NO with graphitic carbon nitride (g-C3N4) activated by the alkaline hydrothermal treatment and mechanism analysis. J. Hazard. Mater. 2015, 300, 598–606. [Google Scholar] [CrossRef] [PubMed]
  76. Sano, T.; Tsutsui, S.; Koike, K.; Hirakawa, T.; Teramoto, Y.; Negishi, N.; Takeuchi, K. Activation of graphitic carbon nitride (g-C3N4) by alkaline hydrothermal treatment for photocatalytic NO oxidation in gas phase. J. Mater. Chem. A 2013, 1, 6489. [Google Scholar] [CrossRef]
  77. Li, Y.; Jin, R.; Xing, Y.; Li, J.; Song, S.; Liu, X.; Li, M.; Jin, R. Macroscopic foam-like holey ultrathin g-C3N4 nanosheets for drastic improvement of visible-light photocatalytic activity. Adv. Energy Mater. 2016, 6, 1601273. [Google Scholar] [CrossRef]
  78. She, X.; Xu, H.; Xu, Y.; Yan, J.; Xia, J.; Xu, L.; Song, Y.; Jiang, Y.; Zhang, Q.; Li, H. Exfoliated graphene-like carbon nitride in organic solvents: Enhanced photocatalytic activity and highly selective and sensitive sensor for the detection of trace amounts of Cu2+. J. Mater. Chem. A 2014, 2, 2563. [Google Scholar] [CrossRef]
  79. Dong, F.; Li, Y.; Wang, Z.; Ho, W.-K. Enhanced visible light photocatalytic activity and oxidation ability of porous graphene-like g-C3N4 nanosheets via thermal exfoliation. Appl. Surf. Sci. 2015, 358, 393–403. [Google Scholar] [CrossRef]
  80. Zhang, Y.; Zong, S.; Cheng, C.; Shi, J.; Guo, P.; Guan, X.; Luo, B.; Shen, S.; Guo, L. Rapid high-temperature treatment on graphitic carbon nitride for excellent photocatalytic H2-evolution performance. Appl. Catal. B Environ. 2018, 233, 80–87. [Google Scholar] [CrossRef]
  81. Zhao, D.; Chen, J.; Dong, C.-L.; Zhou, W.; Huang, Y.-C.; Mao, S.S.; Guo, L.; Shen, S. Interlayer interaction in ultrathin nanosheets of graphitic carbon nitride for efficient photocatalytic hydrogen evolution. J. Catal. 2017, 352, 491–497. [Google Scholar] [CrossRef]
  82. Huangfu, Y.; Zheng, T.; Zhang, K.; She, X.; Xu, H.; Fang, Z.; Xie, K. Facile fabrication of permselective g-C3N4 separator for improved lithium-sulfur batteries. Electrochim. Acta 2018, 272, 60–67. [Google Scholar] [CrossRef]
  83. Hou, Y.; Yang, J.; Lei, C.; Yang, B.; Li, Z.; Xie, Y.; Zhang, X.; Lei, L.; Chen, J. Nitrogen vacancy structure driven photoeletrocatalytic degradation of 4-Chlorophenol using porous graphitic carbon nitride nanosheets. ACS Sustain. Chem. Eng. 2018, 6, 6497–6506. [Google Scholar] [CrossRef]
  84. Liang, Q.; Li, Z.; Huang, Z.-H.; Kang, F.; Yang, Q.-H. Holey graphitic carbon nitride nanosheets with carbon vacancies for highly improved photocatalytic hydrogen production. Adv. Funct. Mater. 2015, 25, 6885–6892. [Google Scholar] [CrossRef]
  85. Song, X.; Yang, Q.; Jiang, X.; Yin, M.; Zhou, L. Porous graphitic carbon nitride nanosheets prepared under self-producing atmosphere for highly improved photocatalytic activity. Appl. Catal. B Environ. 2017, 217, 322–330. [Google Scholar] [CrossRef]
  86. Song, T.; Zhang, P.; Wang, T.; Ali, A.; Zeng, H. Alkali-assisted fabrication of holey carbon nitride nanosheet with tunable conjugated system for efficient visible-light-driven water splitting. Appl. Catal. B Environ. 2018, 224, 877–885. [Google Scholar] [CrossRef]
  87. Hu, P.; Chen, C.; Zeng, R.; Xiang, J.; Huang, Y.; Hou, D.; Li, Q.; Huang, Y. Facile synthesis of bimodal porous graphitic carbon nitride nanosheets as efficient photocatalysts for hydrogen evolution. Nano Energy 2018, 50, 376–382. [Google Scholar] [CrossRef]
  88. Nie, Y.; Zhu, Y.A.; Lu, X.; Qiu, J.; Wang, B.; Xie, Z.; Le, Z. Cu doped crystalline carbon nitride with increased carrier migration efficiency for uranyl photoreduction. Chem. Eng. J. 2023, 477, 146908. [Google Scholar] [CrossRef]
  89. Li, Y.; Xue, Y.; Gao, X.; Wang, L.; Liu, X.; Wang, Z.; Shen, S. Cayanamide group functionalized crystalline carbon nitride aerogel for efficient CO2 photoreduction. Adv. Funct. Mater. 2023, 34, 2312634. [Google Scholar] [CrossRef]
  90. Wang, X.-J.; Tian, X.; Li, F.-T.; Li, Y.-P.; Zhao, J.; Hao, Y.-J.; Liu, Y. Synchronous surface hydroxylation and porous modification of g-C3N4 for enhanced photocatalytic H2 evolution efficiency. Int. J. Hydrogen Energy 2016, 41, 3888–3895. [Google Scholar] [CrossRef]
  91. Zeng, Y.; Li, H.; Luo, J.; Yuan, J.; Wang, L.; Liu, C.; Xia, Y.; Liu, M.; Luo, S.; Cai, T.; et al. Sea-urchin-structure g-C3N4 with narrow bandgap (˜2.0 eV) for efficient overall water splitting under visible light irradiation. Appl. Catal. B Environ. 2019, 249, 275–281. [Google Scholar] [CrossRef]
  92. Li, Y.; Xu, X.; Wang, J.; Luo, W.; Zhang, Z.; Cheng, X.; Wu, J.; Yang, Y.; Chen, G.; Sun, S.; et al. Post-redox engineering electron configurations of atomic thick C3N4 nanosheets for enhanced photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2020, 270, 118855. [Google Scholar] [CrossRef]
  93. Ai, L.; Zhao, Z.; Song, X.; Tang, Y.; Feng, K.; Wang, X.; Zhang, B.; Bi, H.; Qiu, J. Curved-Slit Effect Induced Nucleation for g-C3N4 Nanorings with Double Concaves and Enhanced Photocatalysis. Adv. Funct. Mater. 2025, 35, 2506395. [Google Scholar] [CrossRef]
  94. An, N.; Zhang, X.; Chen, Y.; Wang, Z.; Qiu, J.; Gao, B.; Li, Q. A Self-floating Photothermal/Photocatalytic Evaporator for Simultaneous High-Efficiency Evaporation and Purification of Volatile Organic Wastewater. Adv. Funct. Mater. 2025, 35, 2500777. [Google Scholar] [CrossRef]
  95. Torres-Pinto, A.; Díez, A.M.; Silva, C.G.; Faria, J.L.; Sanromán, M.Á.; Silva, A.M.T.; Pazos, M. Photoelectrocatalytic degradation of pharmaceuticals promoted by a metal-free g-C3N4 catalyst. Chem. Eng. J. 2023, 476, 146761. [Google Scholar] [CrossRef]
  96. He, S.; Wang, X.; Tang, L.; Wang, J.; Chen, J.; Zhang, Y. A Novel G-C3N4 Modified Biogenic Mackinawite Mediated by SRB for Boosting Highly Efficient Adsorption and Catalytic Degradation of Antibiotics in Photo-Fenton Process. Small 2024, 20, 2408723. [Google Scholar] [CrossRef] [PubMed]
  97. Sun, J.; Wang, G. Achieving Near Infrared Photodegradation by the Synergistic Effect of Z-Scheme Heterojunction and Antenna of Rare Earth Single Atoms. Small 2025, 21, 2412148. [Google Scholar] [CrossRef]
  98. Jiang, W.; Luo, W.; Zong, R.; Yao, W.; Li, Z.; Zhu, Y. Polyaniline/carbon nitride nanosheets composite hydrogel: A separation-free and high-efficient photocatalyst with 3D hierarchical structure. Small 2016, 12, 4370–4378. [Google Scholar] [CrossRef]
  99. Zhang, Z.; Pan, Z.; Guo, Y.; Wong, P.K.; Zhou, X.; Bai, R. In-situ growth of all-solid Z-scheme heterojunction photocatalyst of Bi7O9I3/g-C3N4 and high efficient degradation of antibiotic under visible light. Appl. Catal. B Environ. 2020, 261, 118212. [Google Scholar] [CrossRef]
  100. An, S.; Zhang, G.; Wang, T.; Zhang, W.; Li, K.; Song, C.; Miller, J.T.; Miao, S.; Wang, J.; Guo, X. High-density ultra-small clusters and single-atom Fe sites embedded in graphitic carbon nitride (g-C3N4) for hghly efficient catalytic advanced oxidation processes. ACS Nano 2018, 12, 9441–9450. [Google Scholar] [CrossRef]
  101. Gao, M.; Feng, J.; He, F.; Zeng, W.; Wang, X.; Ren, Y.; Wei, T. Carbon microspheres work as an electron bridge for degrading high concentration MB in CoFe2O4@carbon microsphere/g-C3N4 with a hierarchical sandwich-structure. Appl. Surf. Sci. 2020, 507, 145167. [Google Scholar] [CrossRef]
  102. Le, Z.; Xiong, C.; Gong, J.; Wu, X.; Pan, T.; Chen, Z.; Xie, Z. Self-cleaning isotype g-C3N4 heterojunction for efficient photocatalytic reduction of hexavalent uranium under visible light. Environ. Pollut. 2020, 260, 114070. [Google Scholar] [CrossRef]
  103. Wang, X.; Xie, Y.; Chen, X.; Zhou, X.; Hu, W.; Li, P.; Li, Y.; Zhang, Y.; Wang, Y. A novel g-C3N4 modified biosynthetic Fe(III)-hydroxysulfate for efficient photoreduction of Cr(VI) in wastewater treatment under visible light irradiation. Chem. Eng. J. 2020, 398, 125632. [Google Scholar] [CrossRef]
  104. Liao, Q.; Pan, W.; Zou, D.; Shen, R.; Sheng, G.; Li, X.; Zhu, Y.; Dong, L.; Asiri, A.M.; Alamry, K.A.; et al. Using of g-C3N4 nanosheets for the highly efficient scavenging of heavy metals at environmental relevant concentrations. J. Mol. Liq. 2018, 261, 32–40. [Google Scholar] [CrossRef]
  105. Lei, Z.-D.; Xue, Y.-C.; Chen, W.-Q.; Li, L.; Qiu, W.-H.; Zhang, Y.; Tang, L. The influence of carbon nitride nanosheets doping on the crystalline formation of MIL-88B(Fe) and the photocatalytic activities. Small 2018, 14, 1802045. [Google Scholar] [CrossRef]
  106. Chen, S.; Lu, W.; Han, J.; Zhong, H.; Xu, T.; Wang, G.; Chen, W. Robust three-dimensional g-C3N4@cellulose aerogel enhanced by cross-linked polyester fibers for simultaneous removal of hexavalent chromium and antibiotics. Chem. Eng. J. 2019, 359, 119–129. [Google Scholar] [CrossRef]
  107. Wu, J.-H.; Shao, F.-Q.; Luo, X.-Q.; Xu, H.-J.; Wang, A.-J. Pd nanocones supported on g-C3N4: An efficient photocatalyst for boosting catalytic reduction of hexavalent chromium under visible-light irradiation. Appl. Surf. Sci. 2019, 471, 935–942. [Google Scholar] [CrossRef]
  108. Luo, J.; Dong, G.; Zhu, Y.; Yang, Z.; Wang, C. Switching of semiconducting behavior from n-type to p-type induced high photocatalytic NO removal activity in g-C3N4. Appl. Catal. B Environ. 2017, 214, 46–56. [Google Scholar] [CrossRef]
  109. Ou, M.; Wan, S.; Zhong, Q.; Zhang, S.; Song, Y.; Guo, L.; Cai, W.; Xu, Y. Hierarchical Z-scheme photocatalyst of g-C3N4@Ag/BiVO4 (040) with enhanced visible-light-induced photocatalytic oxidation performance. Appl. Catal. B Environ. 2018, 221, 97–107. [Google Scholar] [CrossRef]
  110. Zhou, M.; Dong, G.; Yu, F.; Huang, Y. The deep oxidation of NO was realized by Sr multi-site doped g-C3N4 via photocatalytic method. Appl. Catal. B Environ. 2019, 256, 117825. [Google Scholar] [CrossRef]
  111. Liao, J.; Cui, W.; Li, J.; Sheng, J.; Wang, H.; Dong, X.A.; Chen, P.; Jiang, G.; Wang, Z.; Dong, F. Nitrogen defect structure and NO+ intermediate promoted photocatalytic NO removal on H2 treated g-C3N4. Chem. Eng. J. 2020, 379, 122282. [Google Scholar] [CrossRef]
  112. Hu, J.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. 3D aerogel of graphitic carbon nitride modified with perylene imide and graphene oxide for highly efficient nitric oxide removal under visible light. Small 2018, 14, e1800416. [Google Scholar] [CrossRef]
  113. Li, X.; Dong, G.; Guo, F.; Zhu, P.; Huang, Y.; Wang, C. Enhancement of photocatalytic NO removal activity of g-C3N4 by modification with illite particles. Environ. Sci. Nano 2020, 7, 1990–1998. [Google Scholar] [CrossRef]
  114. dos Santos, G.; Tian, L.; Gonçalves, R.; García, H.; Rossi, L. Boosting CO2 Photoreduction Efficiency of Carbon Nitride via S-scheme g-C3N4/Fe2TiO5 Heterojunction. Adv. Funct. Mater. 2025, 35, 2422055. [Google Scholar] [CrossRef]
  115. Liu, M.; Huang, L.; Su, K.; Liang, J.; Wu, Y.; Luo, P.; Song, H.; Du, L.; Cui, Z. Multiple Ion Rectification Strategy Regulated Polyethylene Glycol-Based Polymer Electrolyte for Stable High-Voltage Lithium Metal Batteries. Adv. Funct. Mater. 2025, 35, 2425417. [Google Scholar] [CrossRef]
  116. Zhu, X.; Zhou, E.; Tai, X.; Zong, H.; Yi, J.; Yuan, Z.; Zhao, X.; Huang, P.; Xu, H.; Jiang, Z. g-C3N4 S-Scheme Homojunction through Van der Waals Interface Regulation by Intrinsic Polymerization Tailoring for Enhanced Photocatalytic H2 Evolution and CO2 Reduction. Angew. Chem. Int. Ed. 2025, 64, e202425439. [Google Scholar] [CrossRef]
  117. Yadav, D.K.; Latiyan, S.; Devan, R.S.; Urkude, R.R.; Rajput, P.; Singh, A.; Deka, S. Breaking Barriers: Synergistic Interactions Between Pt Single Atoms and Nitrogen-Rich g-C3N4 for Maximized Photocatalytic Hydrogen Production. Small 2025, 21, 2503843. [Google Scholar] [CrossRef] [PubMed]
  118. Wang, Y.; Li, H.; Zhai, B.; Li, X.; Niu, P.; Odent, J.; Wang, S.; Li, L. Highly Crystalline Poly(heptazine imide)-Based Carbonaceous Anodes for Ultralong Lifespan and Low-Temperature Sodium-Ion Batteries. ACS Nano 2024, 18, 3456–3467. [Google Scholar] [CrossRef]
  119. Zhuang, C.; Chang, Y.; Li, W.; Li, S.; Xu, P.; Zhang, H.; Zhang, Y.; Zhang, C.; Gao, J.; Chen, G.; et al. Light-Induced Variation of Lithium Coordination Environment in g-C3N4 Nanosheet for Highly Efficient Oxygen Reduction Reactions. ACS Nano 2024, 18, 5206–5217. [Google Scholar] [CrossRef]
  120. Sun, Y.-J.; He, J.-Y.; Zhang, D.; Wang, X.-J.; Zhao, J.; Liu, R.-H.; Li, F.-T. Simultaneous construction of dual-site phosphorus modified g-C3N4 and its synergistic mechanism for enhanced visible-light photocatalytic hydrogen evolution. Appl. Surf. Sci. 2020, 517, 146192. [Google Scholar] [CrossRef]
  121. Zhang, J.; Huang, Y.; Nie, T.; Wang, R.; He, B.; Han, B.; Wang, H.; Tian, Y.; Gong, Y. Enhanced visible-light photocatalytic H2 production of hierarchical g-C3N4 hexagon by one-step self-assembly strategy. Appl. Surf. Sci. 2020, 499, 143942. [Google Scholar] [CrossRef]
  122. Che, W.; Cheng, W.; Yao, T.; Tang, F.; Liu, W.; Su, H.; Huang, Y.; Liu, Q.; Liu, J.; Hu, F.; et al. Fast photoelectron transfer in (Cring)-C3N4 plane heterostructural nanosheets for overall water splitting. J. Am. Chem. Soc. 2017, 139, 3021–3026. [Google Scholar] [CrossRef] [PubMed]
  123. Xiao, X.; Gao, Y.; Zhang, L.; Zhang, J.; Zhang, Q.; Li, Q.; Bao, H.; Zhou, J.; Miao, S.; Chen, N.; et al. A promoted charge separation/transfer system from Cu single atoms and C3N4 layers for efficient photocatalysis. Adv. Mater. 2020, 32, 2003082. [Google Scholar] [CrossRef] [PubMed]
  124. Samanta, S.; Yadav, R.; Kumar, A.; Kumar Sinha, A.; Srivastava, R. Surface modified C, O co-doped polymeric g-C3N4 as an efficient photocatalyst for visible light assisted CO2 reduction and H2O2 production. Appl. Catal. B Environ. 2019, 259, 118054. [Google Scholar] [CrossRef]
  125. Mo, Z.; Zhu, X.; Jiang, Z.; Song, Y.; Liu, D.; Li, H.; Yang, X.; She, Y.; Lei, Y.; Yuan, S.; et al. Porous nitrogen-rich g-C3N4 nanotubes for efficient photocatalytic CO2 reduction. Appl. Catal. B Environ. 2019, 256, 117854. [Google Scholar] [CrossRef]
  126. Li, P.; Liu, L.; An, W.; Wang, H.; Guo, H.; Liang, Y.; Cui, W. Ultrathin porous g-C3N4 nanosheets modified with AuCu alloy nanoparticles and C-C coupling photothermal catalytic reduction of CO2 to ethanol. Appl. Catal. B Environ. 2020, 266, 118618. [Google Scholar] [CrossRef]
  127. Ma, W.; Wang, N.; Guo, Y.; Yang, L.; Lv, M.; Tang, X.; Li, S. Enhanced photoreduction CO2 activity on g-C3N4: By synergistic effect of nitrogen defective-enriched and porous structure, and mechanism insights. Chem. Eng. J. 2020, 388, 124288. [Google Scholar] [CrossRef]
  128. Wang, Y.; Xia, Q.; Bai, X.; Ge, Z.; Yang, Q.; Yin, C.; Kang, S.; Dong, M.; Li, X. Carbothermal activation synthesis of 3D porous g-C3N4/carbon nanosheets composite with superior performance for CO2 photoreduction. Appl. Catal. B Environ. 2018, 239, 196–203. [Google Scholar] [CrossRef]
  129. Cheng, L.; Yin, H.; Cai, C.; Fan, J.; Xiang, Q. Single Ni atoms anchored on porous few-layer g-C3N4 for photocatalytic CO2 reduction: The role of edge confinement. Small 2020, 16, 2002411. [Google Scholar] [CrossRef]
  130. Yao, X.; Zhang, W.; Huang, J.; Du, Z.; Hong, X.; Chen, X.; Hu, X.; Wang, X. Enhanced photocatalytic nitrogen fixation of Ag/B-doped g-C3N4 nanosheets by one-step in-situ decomposition-thermal polymerization method. Appl. Catal. A Gen. 2020, 601, 117647. [Google Scholar] [CrossRef]
  131. Kwon, N.H.; Shin, S.-J.; Jin, X.; Jung, Y.; Hwang, G.-S.; Kim, H.; Hwang, S.-J. Monolayered g-C3N4 nanosheet as an emerging cationic building block for bifunctional 2D superlattice hybrid catalysts with controlled defect structures. Appl. Catal. B Environ. 2020, 277, 119191. [Google Scholar] [CrossRef]
  132. Qiu, P.; Xu, C.; Zhou, N.; Chen, H.; Jiang, F. Metal-free black phosphorus nanosheets-decorated graphitic carbon nitride nanosheets with C P bonds for excellent photocatalytic nitrogen fixation. Appl. Catal. B Environ. 2018, 221, 27–35. [Google Scholar] [CrossRef]
  133. Cao, S.; Fan, B.; Feng, Y.; Chen, H.; Jiang, F.; Wang, X. Sulfur-doped g-C3N4 nanosheets with carbon vacancies: General synthesis and improved activity for simulated solar-light photocatalytic nitrogen fixation. Chem. Eng. J. 2018, 353, 147–156. [Google Scholar] [CrossRef]
  134. Yin, H.; Li, S.-L.; Gan, L.-Y.; Wang, P. Pt-embedded in monolayer g-C3N4 as a promising single-atom electrocatalyst for ammonia synthesis. J. Mater. Chem. A 2019, 7, 11908–11914. [Google Scholar] [CrossRef]
  135. Zhang, J.; Li, J.-Y.; Wang, W.-P.; Zhang, X.-H.; Tan, X.-H.; Chu, W.-G.; Guo, Y.-G. Microemulsion assisted assembly of 3D porous S/Graphene@g-C3N4 hybrid sponge as free-standing cathodes for high energy density Li-S batteries. Adv. Energy Mater. 2018, 8, 1702839. [Google Scholar] [CrossRef]
  136. Tang, Y.; Wang, X.; Chen, J.; Wang, X.; Wang, D.; Mao, Z. Templated transformation of g-C3N4 nanosheets into nitrogen-doped hollow carbon sphere with tunable nitrogen-doping properties for application in Li-ions batteries. Carbon 2020, 168, 458–467. [Google Scholar] [CrossRef]
  137. Wang, M.; Liang, Q.; Han, J.; Tao, Y.; Liu, D.; Zhang, C.; Lv, W.; Yang, Q.-H. Catalyzing polysulfide conversion by g-C3N4 in a graphene network for long-ife lithium-sulfur batteries. Nano Res. 2018, 11, 3480–3489. [Google Scholar] [CrossRef]
  138. Wang, S.; Shi, Y.; Fan, C.; Liu, J.; Li, Y.; Wu, X.-L.; Xie, H.; Zhang, J.; Sun, H. Layered g-C3N4@reduced graphene oxide composites as anodes with improved rate performance for lithium-ion batteries. ACS Appl. Mater. Interfaces 2018, 10, 30330–30336. [Google Scholar] [CrossRef]
  139. Kang, S.; Huang, W.; Zhang, L.; He, M.; Xu, S.; Sun, D.; Jiang, X. Moderate bacterial etching allows scalable and clean delamination of g-C3N4 with enriched unpaired electrons for highly improved photocatalytic water disinfection. ACS Appl. Mater. Interfaces 2018, 10, 13796–13804. [Google Scholar] [CrossRef]
  140. Li, R.; Ren, Y.; Zhao, P.; Wang, J.; Liu, J.; Zhang, Y. Graphitic carbon nitride (g-C3N4) nanosheets functionalized composite membrane with self-cleaning and antibacterial performance. J. Hazard. Mater. 2019, 365, 606–614. [Google Scholar] [CrossRef]
  141. Wang, L.; Zhang, X.; Yu, X.; Gao, F.; Shen, Z.; Zhang, X.; Ge, S.; Liu, J.; Gu, Z.; Chen, C. An all-organic semiconductor C3N4/PDINH heterostructure with advanced antibacterial photocatalytic therapy activity. Adv. Mater. 2019, 31, 1901965. [Google Scholar] [CrossRef] [PubMed]
  142. Saha, D.; Visconti, M.C.; Desipio, M.M.; Thorpe, R. Inactivation of antibiotic resistance gene by ternary nanocomposites of carbon nitride, reduced graphene oxide and iron oxide under visible light. Chem. Eng. J. 2020, 382, 122857. [Google Scholar] [CrossRef]
  143. Chen, S.; Lv, Y.; Shen, Y.; Ji, J.; Zhou, Q.; Liu, S.; Zhang, Y. Highly sensitive and quality self-testable electrochemiluminescence assay of DNA methyltransferase activity using multifunctional sandwich-assembled carbon nitride nanosheets. ACS Appl. Mater. Interfaces 2018, 10, 6887–6894. [Google Scholar] [CrossRef]
  144. Niazi, S.; Khan, I.M.; Yu, Y.; Pasha, I.; Lv, Y.; Mohsin, A.; Mushtaq, B.S.; Wang, Z. A novel fluorescent aptasensor for aflatoxin M1 detection using rolling circle amplification and g-C3N4 as fluorescence quencher. Sens. Actuators B Chem. 2020, 315, 128049. [Google Scholar] [CrossRef]
  145. Yuan, Q.; Li, L.; Tang, Y.; Zhang, X. A facile Pt-doped g-C3N4 photocatalytic biosensor for visual detection of superoxide dismutase in serum samples. Sens. Actuators B Chem. 2020, 318, 128238. [Google Scholar] [CrossRef]
  146. Xu, Y.; Ding, L.; Wen, Z.; Zhang, M.; Jiang, D.; Guo, Y.; Ding, C.; Wang, K. Core-shell LaFeO3@g-C3N4 p-n heterostructure with improved photoelectrochemical performance for fabricating streptomycin aptasensor. Appl. Surf. Sci. 2020, 511, 145571. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of high-SSA g-C3N4 for various applications.
Figure 1. Schematic illustration of high-SSA g-C3N4 for various applications.
Nanomaterials 15 00956 g001
Figure 2. Layer stacking structure and basic units of g-C3N4.
Figure 2. Layer stacking structure and basic units of g-C3N4.
Nanomaterials 15 00956 g002
Figure 5. (a) Schematic illustration of the preparation, photocatalytic H2 production, and photocatalytic degradation of MSCN nanocapsules. Reproduced with permission [53]. Copyright © 2016 American Chemical Society. (b) Schematic illustration of the formation process of tetragonal carbon nitride hollow tubes in the molten salt medium. Reproduced with permission [43]. Copyright © 2018 Elsevier.
Figure 5. (a) Schematic illustration of the preparation, photocatalytic H2 production, and photocatalytic degradation of MSCN nanocapsules. Reproduced with permission [53]. Copyright © 2016 American Chemical Society. (b) Schematic illustration of the formation process of tetragonal carbon nitride hollow tubes in the molten salt medium. Reproduced with permission [43]. Copyright © 2018 Elsevier.
Nanomaterials 15 00956 g005
Figure 8. The characteristics of pre-treatment, in-processing, and post-treatment methods for the preparation of high-SSA g-C3N4.
Figure 8. The characteristics of pre-treatment, in-processing, and post-treatment methods for the preparation of high-SSA g-C3N4.
Nanomaterials 15 00956 g008
Figure 9. Number of articles on precursor treatment strategies from 2016 to 2023.
Figure 9. Number of articles on precursor treatment strategies from 2016 to 2023.
Nanomaterials 15 00956 g009
Figure 10. Precursors and treatment methods commonly used in precursor mixing strategies.
Figure 10. Precursors and treatment methods commonly used in precursor mixing strategies.
Nanomaterials 15 00956 g010
Figure 11. Statistical graph of carbon nitride applications in various fields (numbers of publications using “g-C3N4” or “carbon nitride” as topic keywords since 2016), (a): quantity; (b): percentage of quantity.
Figure 11. Statistical graph of carbon nitride applications in various fields (numbers of publications using “g-C3N4” or “carbon nitride” as topic keywords since 2016), (a): quantity; (b): percentage of quantity.
Nanomaterials 15 00956 g011
Figure 12. Annual distribution map of articles published in the fields of (a) degradation of pollutants in water and (b) energy (numbers of publications using “g-C3N4” or “carbon nitride” as topic keywords since 2016).
Figure 12. Annual distribution map of articles published in the fields of (a) degradation of pollutants in water and (b) energy (numbers of publications using “g-C3N4” or “carbon nitride” as topic keywords since 2016).
Nanomaterials 15 00956 g012
Table 1. Influencing factors and raw materials [20,21,25,26,27,28,29].
Table 1. Influencing factors and raw materials [20,21,25,26,27,28,29].
ReferencesPrecursor 1Precursor 2Mixing ModeCalcining ProcessResults
Bulk g-C3N4 BETIncreased BETMorphologyApplication
[28]Melamine
0.5 g
Urea
5 g
-550 °C for 4 h 5 °C min−1
in a nitrogen
5.6 m2/g42.2 m2/gNanomaterials 15 00956 i001Photocatalytic hydrogen evolution
[21]Cyanuric acid
2.58 g
Melamine
2.52 g
Stirred for 2 h550 °C for 4 h17.74 m2/g81.58 m2/gNanomaterials 15 00956 i002H2 production activity and degradation rate
[25]5 g of formic
acid
3 g of melamineHydrothermal treatment550 °C for 4 h-81.4 m2/gNanomaterials 15 00956 i003Photocatalytic hydrogen evolution
[20]1.0 g melamine2.0 g hydroxylamine
hydrochloride
Hydrothermal process520 °C for 4 h in air3.9 m2/g129.4 m2/gNanomaterials 15 00956 i004Photocatalytic H2O2 production
[26]4 g of melamine50 mL of N,N-
dimethylformamide
Fully mixed at 25 °C for 0.5 h550 °C for 4 h under air11.23 m2/g181.74 m2/gNanomaterials 15 00956 i005Photocatalytic hydrogen evolution
[27]MelamineCyanuric acid
Phosphorous
acid
Hydrothermal process520 °C for 4 h--Nanomaterials 15 00956 i006Electrochemiluminescence
[29]0.01 mol melamine0.01 mol
cyanuric acid
Stirred for 12 h at room temperature550 °C for 4 h 5 °C min−110.83 m2/g130 m2/gNanomaterials 15 00956 i007Photocatalytic overall water splitting
Table 2. The influencing factors and results of precursor pretreatment [32,33,34,35,36,37,38,39,40,42,43].
Table 2. The influencing factors and results of precursor pretreatment [32,33,34,35,36,37,38,39,40,42,43].
ReferencesPrecursorAdded ReagentTreatment ModeCalcining ProcessResults
Bulk g-C3N4 BETIncreased BETMorphologyApplication
[32]Melamine1 mL HCl (37%)In 80 mL of hot distilled water
Stirring for 30 min
500 °C for 2 h 20 °C/min
520 °C for 2 h
8.5 m2/g345 m2/gNanomaterials 15 00956 i008Photocatalytic activity for NO removal
[33]Melamine3 mL concentrated HCl (1:1, v/v)In 30 mL of absolute alcohol
Stirring for 30 min
550 °C for 4 h12.7 m2/g26.2 m2/gNanomaterials 15 00956 i009Photocatalytic degradation
[34]Melamine0.6 M HNO3 solution (50 mL)50 mL of ethylene glycol
Stirring at room temperature
550 °C for 2 h16.6 m2/g86.4 m2/gNanomaterials 15 00956 i010Photocatalytic hydrogen evolution
[35]MelamineH2SO4:H2O
1:1 in volume
In distilled water (100 mL)
Stirring for 2 h
380 °C in 5 min
600 °C for 4 h 1 °C min−1
in Ar
8.6 m2/g15.6 m2/g-Photocatalytic hydrogen evolution
[36]Urea50 mL of methanolDiethyl
until white jellylike crystallization occurred
600 °C for 2 h 2.3 °C min−1
in Ar
43.1 m2/g228.4 m2/gNanomaterials 15 00956 i011Photocatalytic hydrogen evolution
[37]MelamineDried dimethyl sulfoxide 100 mL180 °C under magnetic stirring-7.94 m2/g669.15 m2/gNanomaterials 15 00956 i012Photocatalytic degradation
[42]DicyandiamideNH4ClFrozen in liquid nitrogen550 °C for 4 h 3 °C min−1
in N2
-65 m2/gNanomaterials 15 00956 i013Photocatalytic hydrogen evolution
[43]Melamine352 °C for LiCl-KClMilled together450 °C for 5 h 4 °C min−1
in air
7 m2/g128 m2/gNanomaterials 15 00956 i014Photocatalytic degradation
[38]MelamineDeionized water (40 mL)200 °C for 12 h550 °C for 4 h 2 °C min−18.6 m2/g127.8 m2/gNanomaterials 15 00956 i015Photocatalytic hydrogen evolution
[39]MelamineDeionized water (70 mL)180 °C for 12 h550 °C for 3 h 2.5 °C min−119.9 m2/g67.5 m2/gNanomaterials 15 00956 i016Photocatalytic activity for NO removal
[40]DicyandiamideDeionized water (65 mL)200 °C for 2 h550 °C for 4 h 5 °C min−112.2 m2/g59.8 m2/gNanomaterials 15 00956 i017Photocatalytic hydrogen evolution
Table 3. Influencing factors and results of template methods [44,45,46,47,48,49,50,51,52,53,54,55,56,57,58].
Table 3. Influencing factors and results of template methods [44,45,46,47,48,49,50,51,52,53,54,55,56,57,58].
ReferencesTemplatePrecursorMixing ModeCalcining ProcessRemoving TemplateResults
Bulk g-C3N4 BETIncreased BETMorphologyApplication
[44]SBA-15Ethane diamine CCl4Refluxed and stirred at 90 °C for 6 h600 °C for 5 h 3.0 °C min−1
in a nitrogen
5 wt. % hydrofluoric acid-505 m2/gNanomaterials 15 00956 i018-
[45]SBA-15
10.7 nm
Ethane diamine CCl4Refluxed and stirred at 90 °C for 6 h600 °C for 5 h 3.0 °C min−1
in nitrogen
5 wt. % hydrofluoric acid-830 m2/gNanomaterials 15 00956 i019The Friedel-Crafts acylation of benzene
[46]SBA-15CyanamideStirred for 1 h550 °C for 4 h 2.3 °C min−1NH4HF2 4 M-239 m2/gNanomaterials 15 00956 i020Photocatalytic Hydrogen Evolution
[47]SBA-15Ammonium thiocyanateStirred at 100 °C to remove water550 °C for 2 hNH4HF2 4 M9 m2/g239 m2/gNanomaterials 15 00956 i021Photocatalytic Hydrogen Evolution
SiO2188 m2/gNanomaterials 15 00956 i022
[48]SBA-15Hexamethylene-tetramineStirred at room temperature750 °C
in nitrogen
40% of HF-1116 m2/gNanomaterials 15 00956 i023Dehydrogenation of ethylbenzene to styrene
[49]SBA-15DicyandiamideVaporized at 70 °C550 °C for 3 hNH4HF2 4 M16.7 m2/g50.1 m2/gNanomaterials 15 00956 i024Photocatalytic degradation of fluoroquinolone antibiotics
[50]MCM-22Ethane diamine CCl4Refluxed at 90 °C for 6 h600 °C for 5 h 3.0 °C min−1
in nitrogen
5 wt. % hydrofluoric acidless than 25 m2/g739 m2/gNanomaterials 15 00956 i025-
[51]SiO2CyanamideStirred for 30 min
(CA at 0.01 N HCl and ethanol pH 2,
adding TEOS)
550 °C for 4 h 2.3 °C min−1
in argon
NH4HF2 4 M-131 m2/gNanomaterials 15 00956 i026Photocatalytic Hydrogen Evolution
[52]SiO2
12 nm
CyanamideStirred at 333 K for 12 h823 K for 4 h 2.3 °C min−1 under N2NH4HF2 4 M10 m2/g160 m2/g
228 m2/g
Nanomaterials 15 00956 i027Photocatalytic H2O2 Production
[53]Multishell SiO2 nanospheresCyanamideStirred at 40 °C for 8 h550 °C for 3h under N2Na2CO3 0.3 M-310.7 m2/gNanomaterials 15 00956 i028Photocatalytic Hydrogen Evolution
[54]Chiral mesoporous SiO2 filmsCyanamideSonicated at 55 °C for 4 h550 °C for 4 h 4 °C min−1
in N2
NH4HF2 4 M6.03 m2/g132.26 m2/gNanomaterials 15 00956 i029Photocatalytic Hydrogen Evolution
[55]SiO2 nanotubes with porous shellsCyanamideStirring for 10 min,
separating, and drying
(three times)
550 °C for 4 h
in N2
10% of HF4.6 m2/g135.1 m2/gNanomaterials 15 00956 i030Photocatalytic Hydrogen Evolution
[56]SiO2 microspheresMelamineIn-air CVD method 320 °C for 2 h550 °C for 3 hNH4HF2 4 M10.1 m2/g29.9 m2/gNanomaterials 15 00956 i031Photocatalytic Hydrogen Evolution
[57]SiO2CyanamideStirred at 25 °C for 24 h
NH4OH
Stirred for about 5 min
550 °C for 4 h 1 °C min−1
in N2
NH5F2 4 M-69.1 m2/gNanomaterials 15 00956 i032Photocatalytic Hydrogen Evolution
[58]KCC-1CyanamideSonication at 55 °C for 4 h
(HCl-treated KCC-1)
550 °C for 4 hNH4HF2 4 M9 m2/g160 m2/gNanomaterials 15 00956 i033Photocatalytic Hydrogen Evolution
Table 5. Influencing factors, assistant methods, and results of ultrasonic processing [68,71,72,78].
Table 5. Influencing factors, assistant methods, and results of ultrasonic processing [68,71,72,78].
ReferencesSynthesis of Bulk g-C3N4Ultrasonic ProcessAssistant MethodResults
Mbulk g-C3N4Solvent/VSolventTreated TimeBand GapBETMorphologyApplication
[68]Melamine 600 °C for 2 h 3 °C/min in air0.1 gWater 100 mL16 h-2.64 eV

2.70 eV
-Nanomaterials 15 00956 i042Bioimaging
[71]Commercial g-C3N40.03 gIPA 10 mL10 h-2.35 eV

2.65 eV
384 m2/gNanomaterials 15 00956 i043Hydrogen Evolution
[78]Dicyandiamide 350 °C for 2 h 600 °C for 2 h0.06 g1,3-BUT 25 mL24 h-2.65 eV

2.79 eV
3.3 m2/g

32.54 m2/g
Nanomaterials 15 00956 i044The sensor for trace amounts of Cu2+ determination
Photocatalytic degradation
[72]Melamine
550 °C for 4 h in static air
2.3 °C/min.
0.5 gEthanol–water
150 mL
10 h-2.70 eV

2.79 eV
12.5 m2/g

59.4 m2/g
Nanomaterials 15 00956 i045Photocatalytic degradation
Table 6. Influencing factors, assistant methods, and results of thermal oxidation treatment [69,73,74,77,79,80,81,82,83,84,85,86,87].
Table 6. Influencing factors, assistant methods, and results of thermal oxidation treatment [69,73,74,77,79,80,81,82,83,84,85,86,87].
ReferencesSynthesis of Bulk g-C3N4Thermal Oxidation TreatmentAssistant MethodResults
Mbulk g-C3N4TemperatureTreated TimeBand GapBETMorphologyApplication
[69]Dicyandiamide 550 °C for 4 h in static air 2.3 °C/min0.4 g500 °C
5 °C/min
2 h
in static air
-2.77 eV

2.97 eV
50 m2/g

306 m2/g
Nanomaterials 15 00956 i046Photocatalytic hydrogen evolution
[79]Thiourea
550 °C for 2 h 15 °C/min
-550 °C2 h
in air
-2.42 eV

2.86 eV
27 m2/g

151 m2/g
Nanomaterials 15 00956 i047Visible light photocatalytic removal of NOx
[80]Melamine 520 °C for 4 h 5 °C min−1
in static air
1.0 g800 °C
600 °C min−1
15 mincooled by circulation cooling water (15 °C)2.64 eV

2.81 eV
7.38 m2/g

60.51 m2/g
Nanomaterials 15 00956 i048Photocatalytic hydrogen evolution
[73]Dicyandiamide 550 °C for 4 h 2 °C/min
in static air
0.5 g500 °C
5 °C/min
2 hput quickly into liquid nitrogen2.68 eV

2.57 eV
15.6 m2/g

142.8 m2/g
Nanomaterials 15 00956 i049Photocatalytic degradation
[74]Urea
550 °C for 9 h 10 °C/min
in air
---A single continuous heating process-40.22 m2/g

117.27 m2/g
Nanomaterials 15 00956 i050Photocatalytic hydrogen evolution
[81]Melamine
520 °C for 4 h in air
5 °C/min
0.5 g520 °C
5 °C/min
4 h
in air
0.5 g
200 mL water
ultrasonication
2 h
2.64 eV

2.72 eV
10.94 m2/g

99.73 m2/g
Nanomaterials 15 00956 i051Photocatalytic hydrogen evolution
[77]Melamine
773 K for 2 h 2 K min −1
793 K for 2 h
1 g793 K
2 K min −1
6 hIncreased calcination time2.67 eV

2.81 eV
10.89 m2/g

277.98 m2/g
Nanomaterials 15 00956 i052Photocatalytic hydrogen evolution
[82]Melamine
550 °C for 4 h 2 °C min−1
0.4 g550 °C30 minwice---Lithium-sulfur batteries
[83]Dicyandiamide 550 °C for 4 h 5 °C min−1
in air
-600 °C2 hH2 atmosphere2.78 eV

1.82 eV
7 m2/g

114 m2/g
Nanomaterials 15 00956 i053Photoeletrocatalytic Degradation of 4-Chlorophenol
[84]Dicyandiamide
550 °C for 4 h
0.3 g510 °C1 hNH3 atmosphere2.59 eV

2.90 eV
6 m2/g

196 m2/g
Nanomaterials 15 00956 i054Photocatalytic hydrogen evolution
[83]Melamine
550 °C for 4 h
5 °C/min
in N2 gas
-300 °C-Self-producted NH3 atmosphere2.78 eV

3.00 eV
6.57 m2/g

38.51 m2/g
Nanomaterials 15 00956 i055Photocatalytic degradation
[86]Melamine
550 °C for 4 h 10 °C min−1
6 g300 °C
2 °C min−1
1 hMass ratio of CN (B): KOH is 1:2 in 50 mL H2O2.55 eV

2.66 eV
10.3 m2/g

265.2 m2/g
Nanomaterials 15 00956 i056Visible-light-driven water splitting
[87]Melamine
550 °C for 2 h in static air
5 °C min−1
0.1 g350 °C
5 °C min−1
1.5 h
in static air
0.1 g of bulk g-C3N4, 0.2 g of KOH and 5 mL of H2O2.53 eV

2.75 eV
219 m2/gNanomaterials 15 00956 i057Photocatalytic hydrogen evolution
Table 7. Influencing factors, assistant methods, and results of post-hydrothermal preparation methods [70,75,76,90,91,92].
Table 7. Influencing factors, assistant methods, and results of post-hydrothermal preparation methods [70,75,76,90,91,92].
ReferencesSynthesis of Bulk g-C3N4Post-HydrothermalAssistant MethodResults
Mbulk g-C3N4Type of SolutionTreated TemperatureTreated TimeBand GapBETMorphologyApplication
[75]Dicyandiamide 550 °C for 4 h 2.9 °C min−11.0 g90 mL NaOH
0.12 M
120 °C18 h-2.75 eV

2.67 eV
29.7 m2/g ↓
64.7 m2/g
Nanomaterials 15 00956 i058Photocatalytic oxidation of gaseous NO
[76]Melamine
550 °C for 2 h
10 °C min−1
1.0 g90 mL NaOH
0.1 M
130 °C18 h--7.7 m2/g

53.7 m2/g
Nanomaterials 15 00956 i059Photocatalytic NO oxidation in gas phase
[70]Dicyandiamide 550 °C for 4 h 4 °C min−1
in air
0.2 g(50-X) mL distilled water -ammonium hydroxides120 °C12 h-2.79 eV

2.91 eV
7.43 m2/g ↓
42.78 m2/g
Nanomaterials 15 00956 i060Photocatalytic hydrogen generation
[90]Melamine
550 °C for 3 h
10 °C/min
0.5 g35 mL ammonium hydroxide (mass fraction = 5%)160 °C4 h-2.76 eV

2.86 eV
14.6 m2/g ↓
44.8 m2/g
Nanomaterials 15 00956 i061Photocatalytic hydrogen generation
[91]Melamine
550 °C for 4 h
2.5 °C/min
in N2
0.23 g60 mL distilled water180 °C6 hPorous g-C3N4
---sealed condensation
2.68 eV

2.07 eV
1.59 m2/g ↓
65.6 m2/g
Nanomaterials 15 00956 i062Overall water splitting
[92]Polycondensation of urea0.5 g10 mL
0.1 M
KOH
150 °C12 hCarbon thermal reduction2.72 eV

2.57 eV
38.7 m2/g ↓
197.0 m2/g
Nanomaterials 15 00956 i063Photocatalytic hydrogen evolution
Table 8. Application of g-C3N4 systems for environmental pollutant mitigation [21,32,33,37,39,43,49,62,73,79,83,93,94,95,96,97].
Table 8. Application of g-C3N4 systems for environmental pollutant mitigation [21,32,33,37,39,43,49,62,73,79,83,93,94,95,96,97].
ReferencesMaterialsMorphologySynthesisBETApplication
[21]g-C3N4Nanomaterials 15 00956 i064Pre-treatment81.58 m2/gH2 production activity and degradation rate
[32]g-C3N4Nanomaterials 15 00956 i065Pre-treatment345 m2/gPhotocatalytic activity for NO removal
[33]g-C3N4Nanomaterials 15 00956 i066Pre-treatment26.2 m2/gPhotocatalytic degradation
[37]g-C3N4Nanomaterials 15 00956 i067Pre-treatment669.15 m2/gPhotocatalytic degradation
[43]g-C3N4Nanomaterials 15 00956 i068In-process128 m2/gPhotocatalytic degradation
[39]g-C3N4Nanomaterials 15 00956 i069Pre-treatment67.5 m2/gPhotocatalytic activity for NO removal
[49]g-C3N4Nanomaterials 15 00956 i070In-process50.1 m2/gPhotocatalytic degradation of fluoroquinolone antibiotics
[62]g-C3N4Nanomaterials 15 00956 i071Post-treatment25.7 m2/gPhoto-reduction of p-nitrophenol
[79]g-C3N4Nanomaterials 15 00956 i072Post-treatment151 m2/gVisible light photocatalytic removal of NOx
[73]g-C3N4Nanomaterials 15 00956 i073Post-treatment142.8 m2/gPhotocatalytic degradation
[83]g-C3N4Nanomaterials 15 00956 i074Post-treatment114 m2/gPhotoeletrocatalytic Degradation of 4-Chlorophenol
[93]g-C3N4Nanomaterials 15 00956 i075In-process241.4 m2/gPhotocatalytic degradation
[94]Mn@g-C3N4
/PANI/wood-derived carbon
Nanomaterials 15 00956 i076Composite materials-Photocatalytic degradation
[95]g-C3N4-Pre-treatment-Photoelectrocatalytic degradation
[96]g-C3N4@biogenic FeSNanomaterials 15 00956 i077Composite materials-Photocatalytic degradation
[97]NaYF4@g-C3N4Nanomaterials 15 00956 i078Composite materials14.10 m2/gPhotocatalytic degradation
Table 9. The application of g-C3N4 nanostructures in energy conversion and storage [25,26,28,36,38,46,51,53,54,55,58,60,65,69,71,74,77,81,82,84,114,115,116,117,118,119].
Table 9. The application of g-C3N4 nanostructures in energy conversion and storage [25,26,28,36,38,46,51,53,54,55,58,60,65,69,71,74,77,81,82,84,114,115,116,117,118,119].
ReferencesMaterialsMorphologySynthesisBETApplication
[28]g-C3N4Nanomaterials 15 00956 i079Pre-treatment42.2 m2/gPhotocatalytic hydrogen evolution
[25]g-C3N4Nanomaterials 15 00956 i080Pre-treatment81.4 m2/gPhotocatalytic hydrogen evolution
[26]g-C3N4Nanomaterials 15 00956 i081Pre-treatment181.74 m2/gPhotocatalytic hydrogen evolution
[36]g-C3N4Nanomaterials 15 00956 i082Pre-treatment228.4 m2/gPhotocatalytic hydrogen evolution
[38]g-C3N4Nanomaterials 15 00956 i083Pre-treatment127.8 m2/gPhotocatalytic hydrogen evolution
[46]g-C3N4Nanomaterials 15 00956 i084In-process239 m2/gPhotocatalytic hydrogen evolution
[51]g-C3N4Nanomaterials 15 00956 i085In-process131 m2/gPhotocatalytic hydrogen evolution
[53]g-C3N4Nanomaterials 15 00956 i086In-process310.7 m2/gPhotocatalytic hydrogen evolution
[54]g-C3N4Nanomaterials 15 00956 i087In-process132.26 m2/gPhotocatalytic hydrogen evolution
[55]g-C3N4Nanomaterials 15 00956 i088In-process135.1 m2/gPhotocatalytic hydrogen evolution
[58]g-C3N4Nanomaterials 15 00956 i089In-process160 m2/gPhotocatalytic hydrogen evolution
[60]g-C3N4Nanomaterials 15 00956 i090Post-treatment55.4 m2/gPhotocatalytic hydrogen evolution and CO2 conversion
[65]g-C3N4Nanomaterials 15 00956 i091Post-treatment205.8 m2/gPhotocatalytic hydrogen evolution
[71]g-C3N4Nanomaterials 15 00956 i092Post-treatment384 m2/gHydrogen evolution
[69]g-C3N4Nanomaterials 15 00956 i093Post-treatment306 m2/gPhotocatalytic hydrogen evolution
[74]g-C3N4Nanomaterials 15 00956 i094Post-treatment117.27 m2/gPhotocatalytic hydrogen evolution
[81]g-C3N4Nanomaterials 15 00956 i095Post-treatment99.73 m2/gPhotocatalytic hydrogen evolution
[77]g-C3N4Nanomaterials 15 00956 i096Post-treatment277.98 m2/gPhotocatalytic hydrogen evolution
[82]g-C3N4-Post-treatment-Lithium–sulfur batteries
[84]g-C3N4Nanomaterials 15 00956 i097Post-treatment196 m2/gPhotocatalytic hydrogen evolution
[114]g-C3N4/Fe2TiO5Nanomaterials 15 00956 i098Composite materials20.28 m2/gCO2 Photoreduction
[115]CN-Nv-C3N4Nanomaterials 15 00956 i099Post-treatment-Lithium metal batteries
[116]g-C3N4 S-Scheme HomojunctionNanomaterials 15 00956 i100Post-treatment122.04 m2/gCO2 photoreduction
[117]Pt NP decorated C3N4Nanomaterials 15 00956 i101Composite materials106.19 m2/gPhotocatalytic hydrogen evolution
[118]CK-CNCNanomaterials 15 00956 i102In-process158.65 m2/gLow-temperature sodium-ion
batteries
[119]Li-C3N4Nanomaterials 15 00956 i103In-process-Oxygen
reduction reactions
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

Gao, M.; Zhao, M.; Yang, Q.; Bao, L.; Chen, L.; Liu, W.; Feng, J. A Review on Pre-, In-Process, and Post-Synthetic Strategies to Break the Surface Area Barrier in g-C3N4 for Energy Conversion and Environmental Remediation. Nanomaterials 2025, 15, 956. https://doi.org/10.3390/nano15130956

AMA Style

Gao M, Zhao M, Yang Q, Bao L, Chen L, Liu W, Feng J. A Review on Pre-, In-Process, and Post-Synthetic Strategies to Break the Surface Area Barrier in g-C3N4 for Energy Conversion and Environmental Remediation. Nanomaterials. 2025; 15(13):956. https://doi.org/10.3390/nano15130956

Chicago/Turabian Style

Gao, Mingming, Minghao Zhao, Qianqian Yang, Lan Bao, Liwei Chen, Wei Liu, and Jing Feng. 2025. "A Review on Pre-, In-Process, and Post-Synthetic Strategies to Break the Surface Area Barrier in g-C3N4 for Energy Conversion and Environmental Remediation" Nanomaterials 15, no. 13: 956. https://doi.org/10.3390/nano15130956

APA Style

Gao, M., Zhao, M., Yang, Q., Bao, L., Chen, L., Liu, W., & Feng, J. (2025). A Review on Pre-, In-Process, and Post-Synthetic Strategies to Break the Surface Area Barrier in g-C3N4 for Energy Conversion and Environmental Remediation. Nanomaterials, 15(13), 956. https://doi.org/10.3390/nano15130956

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