Next Article in Journal
Ce/Mn Co-Doping Induces Synergistic Effects for Low-Temperature NH3-SCR over Ba2Ti5O12 Catalysts
Previous Article in Journal
Applications of Quantum Dots in Photo-Based Advanced Oxidation Processes for the Degradation of Contaminants of Emerging Concern—A Review
Previous Article in Special Issue
Photothermal-Assisted Photocatalytic Degradation of Antibiotic by Black g-C3N4 Materials Derived from C/N Precursors and Tetrachlorofluorescein
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Graphitic Carbon Nitride-Based S-Scheme Heterojunctions: Recent Advances in Photocatalytic Dye Degradation

1
Department of Biological Engineering, Xianning Vocational Technical College, Xianning 437100, China
2
Hubei Province Etnan Special Agricultural Industry Technology Research Institute, Xianning 437100, China
3
Xianning Public Inspection and Testing Center, Xianning 437100, China
4
Hubei Key Laboratory of Radiation Chemistry, School of Nuclear Technology and Chemistry & Biology, and Functional Materials, Hubei University of Science and Technology, Xianning 437100, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(6), 592; https://doi.org/10.3390/catal15060592
Submission received: 14 April 2025 / Revised: 6 June 2025 / Accepted: 11 June 2025 / Published: 15 June 2025
(This article belongs to the Special Issue Advances in Photocatalytic Degradation of Pollutants in Wastewater)

Abstract

With the rapid advancement of industrialization, dye-containing wastewater has emerged as one of the primary pollution sources in aquatic environments, posing a significant threat to ecosystems and human health. S-scheme heterojunction photocatalysis technology, known for its high efficiency and environmental compatibility, is considered a strategic solution for addressing environmental pollution challenges. In recent years, significant progress has been made in the development of S-scheme heterojunction photocatalysts based on graphitic carbon nitride (g-C3N4). However, systematic summaries and in-depth analyses of these advancements remain limited. This study provides a comprehensive review of the research progress of g-C3N4-based S-scheme heterojunction systems in the field of photocatalytic dye degradation. It elaborates on the fundamental concepts, operational mechanisms, and representative applications of these systems while exploring the latest advancements in synthesis strategies, catalytic performance optimization, and the underlying mechanisms. Finally, this review discusses the existing challenges and future prospects of g-C3N4-based S-scheme heterojunction photocatalytic materials, aiming to offer valuable insights and guidance for further research in this area.

Graphical Abstract

1. Introduction

In the contemporary era of industrialization, the environmental pollution issues arising from agricultural modernization and large-scale industrial expansion have posed significant challenges to ecological security and public health. Research has demonstrated that organic pollutants, including surfactants, pesticide residues, pharmaceutical intermediates, and synthetic dyes discharged during production processes, pose substantial risks to the environment [1]. Notably, synthetic dyes released by industries such as textile printing and dyeing, food processing, and pulp and paper manufacturing have become critical concerns in water environmental governance due to their high chemical stability, toxicity, and low biodegradability [2]. To address the treatment of these pollutants, researchers have explored various methods, including adsorption [3], membrane separation technology [4], microbial degradation [5], electrochemical oxidation [6], and photocatalytic degradation. Among these approaches, semiconductor photocatalytic systems leverage the photoluminescence effect of photosensitive materials (e.g., TiO2 and g-C3N4) to generate reactive oxygen species, such as hydroxyl radicals (•OH) and superoxide radicals (•O2⁻), achieving pollutant mineralization and energy transformation [7]. These systems exhibit notable advantages, such as mild reaction conditions, no secondary pollution, and solar-driven operation, offering broad prospects for addressing pollution caused by organic dyes [8]. However, most traditional semiconductor photocatalysts suffer from limitations, including the low efficiency of photogenerated carrier separation, a limited surface area, a narrow absorption spectrum, and poor cycling stability, which significantly impede their practical application [9]. For instance, TiO2 exhibits a wide band gap (approximately 3.2 eV) and low visible light utilization, restricting its photocatalytic activity primarily to the ultraviolet region, which constitutes only about 5% of solar radiation, thereby greatly limiting its effective utilization of sunlight in photocatalysis [10]. Consequently, this has prompted the academic community to focus research efforts on constructing novel photocatalytic systems, particularly catalytic materials with wide spectral response capabilities, which represent a key breakthrough in achieving effective environmental remediation [11]. This review systematically examines the application of novel g-C3N4-based S-scheme heterojunction photocatalytic materials in dye degradation, as illustrated in Figure 1, aiming to provide valuable insights and references for the photocatalytic treatment of organic dyes.

2. S-Scheme Heterojunctions: Proposal and Principles

2.1. Proposal of S-Scheme Heterojunctions

Photocatalytic heterojunction systems are functional structures that effectively integrate two or more semiconductor materials (or semiconductor–metal composites) through heterojunction interface engineering. The primary objective is to significantly enhance photocatalyst performance by leveraging the interfacial synergy between different materials. These heterojunctions, constructed based on energy level alignment principles, are widely recognized in academia as a critical strategy for overcoming the application limitations of photocatalytic technology due to their ability to markedly improve carrier separation efficiency and hole transport capability [12]. They represent an effective and promising approach to enhancing the separation efficiency of photogenerated charges and the oxidation capacity of holes. Current research focuses on developing economically viable heterojunction construction strategies with high carrier separation efficiency, and related advancements have become a major research hotspot in materials science [13]. Based on the relative positions of band gaps, photocatalytic heterojunction systems primarily include type-I, type-II, p-n-type, Z-scheme, and S-scheme configurations [14,15,16]. While traditional type-II heterojunctions achieve spatial carrier separation, they result in a significant attenuation of redox potentials. In contrast, the direct Z-scheme system, inspired by natural photosynthesis, constructs a directional charge transfer channel to effectively maintain the strong redox properties of photogenerated holes and electrons [17]. To address the inherent contradiction between traditional type-II and Z-scheme systems in balancing charge separation and redox capabilities, Yu et al. [17] proposed the stepped (S-scheme) heterojunction theory. This configuration optimizes carrier migration pathways and spatially decouples redox sites by precisely regulating the built-in electric field driven by Fermi level differences, demonstrating superior catalytic performance compared to conventional heterojunction systems [17,18,19]. Studies have confirmed that S-scheme heterojunctions represent a significant breakthrough in overcoming the potential loss of type-II systems and the light absorption limitations of Z-scheme systems [20,21,22].

2.2. Basic Principles of S-Scheme Heterojunctions

S-scheme semiconductor heterostructures are specialized energy band systems formed through the interface coupling of two semiconductor materials with distinct bandgap characteristics, exhibiting an asymmetric S-scheme band-bending distribution at the interface. In contrast to the linear energy band alignment of conventional heterojunctions, this structure generates a dual-bending band configuration during Fermi level equilibration, thereby inducing a unique interfacial charge transfer mechanism. Studies have demonstrated [23] that when high-electron-affinity materials are combined with low-bandgap semiconductor heterostructure components, such S-scheme heterojunctions readily form at the interface. The key advantage of this structure lies in its ability to achieve the efficient separation of photogenerated electron–hole pairs via band gradient engineering while optimizing carrier migration pathways to minimize recombination losses. This dynamic characteristic renders it an ideal platform for constructing advanced optoelectronic devices, particularly in applications such as solar cells, photocatalytic reaction systems, and high-sensitivity photodetectors, showcasing significant application potential [24]. Consequently, the S-scheme charge transfer mechanism represents a rational approach to enhancing photocatalytic performance. Typically, an S-scheme heterojunction can be formed by coupling two semiconductors, where one exhibits a lower Fermi level and functions as an oxidative photocatalyst, while the other possesses a higher Fermi level and serves as a reductive photocatalyst. In S-scheme heterojunctions, the built-in electric field between the semiconductors and the band bending at the interface facilitates interfacial charge transfer and enhances photocatalytic activity [25,26,27]. For instance, when RCN (g-C3N4) and VO (V2O5) are coupled without light exposure, electrons spontaneously transfer from RCN to VO due to the higher Fermi level (EF) of RCN compared to that of VO, as illustrated in Figure 2a [18]. Furthermore, under illumination, the band bending and resultant built-in electric field interact with Coulombic forces, enabling the rapid transfer of excited electrons from the conduction band (CB) of VO across the interface to recombine with holes (h⁺) in the valence band (VB) of RCN. Consequently, electrons accumulate in the CB of RCN, while holes remain in the VB of VO. These retained carriers enable the reduction of O2 to •O2⁻ and the oxidation of adsorbed H2O to •OH, respectively. Ultimately, these highly reactive radicals contribute to the degradation of tetracycline.
In summary, the S-scheme heterojunction system achieves a synergistic enhancement in the separation efficiency of photogenerated charges and redox potentials through a precise band structure design and carrier pathway optimization, significantly improving photocatalytic performance. Therefore, S-scheme heterojunctions represent one of the most promising candidates for photocatalytic applications.

3. g-C3N4’s Role in S-Scheme Heterojunctions and Typical Systems

3.1. g-C3N4’s Role in S-Scheme Heterojunctions

Graphitic carbon nitride, as a novel non-metallic polymer semiconductor, features unique two-dimensional conjugated structures that confer it with remarkable photoelectrochemical properties, including a wide spectral response (visible light absorption edge at approximately 460 nm), an appropriate bandgap structure (~2.7 eV), a high specific surface area, and excellent chemical/thermal stability. These advantages have positioned g-C3N4 as a material of significant interest in environmental catalysis and clean energy applications [30,31]. Studies have demonstrated its substantial potential in photo-driven pollutant mineralization [32] and photoelectrochemical water splitting for hydrogen production [33,34]. However, practical applications of g-C3N4 are still constrained by certain limitations that hinder its photocatalytic efficiency, such as rapid carrier recombination and insufficient hole oxidation potential, which lead to sluggish surface reaction kinetics [35]. To address these challenges, heterojunction construction based on band engineering strategies has emerged as an effective approach to overcoming the performance limitations of single-component materials. By precisely aligning the band structure parameters of the semiconductor components, the interfacial carrier transport dynamics can be optimally regulated [36,37]. Of particular significance is the g-C3N4-based S-scheme heterojunction system, which innovatively achieves the spatial separation of photogenerated electron–hole pairs while effectively preserving the system’s high redox potential through cooperative regulation of the interfacial Schottky barrier and built-in electric field. This synergy of charge separation and retention significantly enhances the catalytic efficiency of surface active sites, offering a new paradigm for overcoming the performance bottlenecks of traditional photocatalytic materials [38].

3.2. Typical Systems of g-C3N4-Based S-Scheme Heterojunctions

In recent years, innovative research on g-C3N4-based S-scheme heterojunction photocatalytic systems has been continuously emerging, with improved optimization strategies and more in-depth exploration of their performance mechanisms. For example, Ha et al. [39] prepared MoS2/g-C3N4 heterojunctions via an in situ synthesis method and constructed an S-scheme charge transfer channel through interface engineering strategies. This catalyst demonstrated excellent photocatalytic performance in Rhodamine B (RhB) degradation, attributed to the directional carrier migration mechanism induced by band alignment. Li et al. [40] reported the main characteristics, design concepts, and photocatalytic applications of g-C3N4-based S-scheme heterojunction photocatalysts. Li et al. [41] successfully achieved RhB photocatalytic degradation on the surface of a novel S-scheme 2D/2D Bi2MoO6/g-C3N4 composite by loading Au as an auxiliary catalyst. Regarding wide-spectrum response regulation, Liu et al. [42] confirmed that the CdS/g-C3N4 system achieved effective carrier separation through an S-scheme mechanism while simultaneously expanding the spectral response range. Guo et al. [43] successfully fabricated an Ag-doped g-C3N4/iron vanadate (FeVO4) S-type heterojunction. The incorporation of Ag not only strengthened the interfacial electric field within the S-type heterojunction but also modulated the production of active species in the Ag/g-C3N4/FeVO4/H2O2/light system. This non-radical photogenerated hole-based mechanism enables highly efficient degradation of tetracycline (TC), with a notable reduction in the toxicity of the degradation intermediates. Le et al. [28] reported the high efficiency of amoxicillin removal by V2O5-g-C3N4 S-scheme nanocomposites under sunlight, as shown in Figure 2b. Figure 2c illustrates the preparation process of vanadium dioxide/carbon nanotube network-structured nanocomposites using a high-temperature calcination method. Dai et al. [44] reported a g-C3N4/ZnFe2O4 S-scheme photocatalyst that was able to remove uranium (VI). Chu et al. [45] constructed a two-dimensional CeO2/g-C3N4 heterointerface via a molten salt template method, which exhibited excellent photocatalytic degradation performance for methylene blue (MB) and tetracycline, attributed to the synergistic effect of the S-scheme mechanism and surface oxygen vacancies. Chen et al. [46] further revealed details of the regulation of methyl orange degradation pathways by the interfacial bonding mode in the C-O chemically bridged CeO2/g-C3N4 S-scheme heterojunction system. In the development of new hybrid systems, Chen et al. [47] designed BiOX/GaMOF heterojunctions, enhancing the removal efficiency of ciprofloxacin to 94.2% through an adsorption–photocatalytic coupling mechanism. The performance advantage originated from the synergy between the porous structure characteristics of metal–organic frameworks and the S-scheme charge transfer mechanism.

4. Application of Graphitic Carbon Nitride-Based S-Scheme Heterojunctions in Dye Degradation

4.1. Application of Graphitic Carbon Nitride-Based S-Scheme Heterojunctions in Cationic Dye Degradation

In recent years, g-C3N4-based S-scheme heterojunctions have exhibited remarkable advantages in the degradation of cationic dyes, attributed to their unique charge separation mechanism and controllable band structure. In contrast to traditional heterojunctions, the S-scheme system achieves a significant enhancement in photocatalytic performance by effectively suppressing the recombination of photogenerated electron–hole pairs through the interfacial built-in electric field and directional carrier migration pathways while preserving the high redox capabilities of each component. By studying methylene blue, Rhodamine B, crystal violet (CV), malachite green (MG), and other target pollutants as examples, researchers have developed highly efficient and stable photocatalytic systems via strategies such as multi-component composites (e.g., GO/g-C3N4/TiO2), plasmonic resonance effects (e.g., Ag and Co nanoparticles), and three-dimensional porous structure designs. These systems typically achieve degradation efficiencies exceeding 95.0%, with some even attaining complete degradation at 100.0%. Mechanistic investigations have revealed that interface band engineering and structural synergy effects are the primary drivers of performance optimization. Specifically, S-scheme heterojunctions establish directional charge transfer channels by precisely aligning energy level positions, while three-dimensional porous frameworks and defect engineering significantly enhance light-harvesting capabilities and reaction kinetics efficiency. This section provides a systematic review of the latest advancements in g-C3N4-based S-scheme heterojunctions for the degradation of cationic dyes, emphasizing interface construction strategies, charge transfer mechanisms, and multi-component synergistic enhancement principles, thereby offering technical insights for the development of advanced environmental remediation materials.

4.1.1. Methylene Blue

Ren et al. [29] innovatively developed a ternary S-scheme heterojunction interface system composed of graphene oxide/graphitic carbon nitride/titanium dioxide through a multi-stage temperature-controlled self-assembly strategy. This study utilized a gradient calcination process to control the interlayer spacing of graphitic carbon nitride and precisely embedded anatase TiO2 nanocrystals into the two-dimensional heterojunction interface of graphitic carbon nitride and graphene oxide via an in situ crystallization strategy (Figure 2d). High-resolution transmission electron microscopy (HRTEM) characterization (Figure 2e) revealed that the composite material exhibited a typical sandwich structure: the graphitic carbon nitride crystal planes and the amorphous carbon layers of graphene oxide formed interlayer channels through π-π stacking, while the anatase TiO2 crystal planes were uniformly distributed on the surface of the two-dimensional substrate (Figure 2f). These results confirmed the successful construction of a ternary heterojunction, indicating that the ternary system was not a simple physical mixture. The GO/g-C3N4/TiO2 composite material achieved a degradation rate of 98.8% for methylene blue after 240 min of visible light irradiation, as shown in Table 1. The performance improvement mechanism can be attributed to the synergistic effect of the electron bridging effect of GO and the precise band alignment of the ternary system, endowing it with excellent photocatalytic activity. Sun et al. [48] developed the Ag/AgI/N-CT-10 plasmonic S-scheme heterojunction system through a multi-step coupling synthesis strategy (hydrothermal–high-temperature calcination–in situ photoreduction). This system utilized nitrogen-vacancy-modified g-C3N4 hollow microtubes (N-CT-10), regulated by potassium hydroxide, as the carrier and achieved the precise loading of Ag/AgI nanoparticles via surface functionalization/modification (Figure 3a). Scanning electron microscopy (SEM) characterization revealed that the composite material exhibited a typical hollow tubular morphology with a significantly increased surface roughness of the tube wall, attributed to the uniform distribution of Ag/AgI hetero-particles. HRTEM analysis (Figure 3b,c) confirmed the presence of clear lattice fringes at the interface: the g-C3N4 (002) plane and the AgI (002) plane formed an interlaced arrangement, while the metallic characteristics of the Ag (110) plane provided a structural basis for the surface plasmon resonance (SPR) effect [49,50,51]. Electron paramagnetic resonance (EPR) spectra (Figure 3d) revealed a significant signal peak at g = 2.005 and, combined with XPS valence-band spectrum shift analysis, confirmed the existence of nitrogen vacancies [52,53]. Under simulated solar light irradiation for 50 min, this catalyst achieved a 96.9% degradation rate for methylene blue. The performance improvement mechanism was attributed to the synergistic effect of the interface charge trap effect induced by nitrogen vacancies and the SPR effect of Ag nanoparticles. Three-dimensional charge density reconstruction analysis based on density functional theory (Figure 3e) demonstrated that the Ag/AgI40%/N-CT-10 system exhibited a significant charge redistribution along the z-axis direction, confirming the formation of a bipolar charge layer structure at the interface. Differential charge density distribution (Figure 3f) quantitatively demonstrated that electrons accumulated on the surfaces of AgI and Ag. This asymmetric charge distribution formed an interface potential gradient, driving photogenerated electrons to migrate along the S-scheme pathway [54]. Therefore, under the premise of following the S-scheme heterojunction charge transfer mechanism, Ag/AgI40%/N-CT-10 exhibited highly efficient carrier separation efficiency, significantly enhancing photocatalytic degradation efficiency [54].
In the development of multi-component composite systems, Leelavathi et al. [27] constructed a g-C3N4/Co/ZnO S-scheme heterostructure via an ultrasonic cavitation-assisted sol–gel method. In this system, Co nanoparticles extended the material’s light response range through the plasmon resonance effect. Under visible light irradiation for 120 min, the g-C3N4/Co/ZnO photocatalyst achieved a 96.3% degradation rate for MB. G. Venkatesh et al. [55] successfully fabricated perovskite-type CaSnO3/g-C3N4 S-scheme heterojunction nanocomposites via a straightforward solid-state route. The CaSnO3/g-C3N4 photocatalytic material achieved a 95.0% degradation efficiency for MB under UV–visible light irradiation for 120 min. A robust solid–solid contact interface was formed between g-C3N4 and CaSnO3, providing additional transport pathways and effectively enhancing the separation of photogenerated electrons and holes. Patra et al. [56] synthesized a PbTiO3/g-C3N4 S-scheme heterojunction through an ultrasonic-assisted hydrothermal method. The PbTiO3/g-C3N4 photocatalyst achieved a 99.4% degradation rate for MB under simulated solar irradiation for 90 min. This hybrid heterojunction exhibited an S-scheme charge transfer mechanism. The incorporation of PbTiO3 nanosheets into the g-C3N4 matrix generated a strong electric field, leading to superior photocatalytic activity. This enhancement can be attributed to the effective separation of photogenerated charge carriers. Abedini et al. [57] synthesized Sm2CeMnO6/g-C3N4 composites via a sol–gel method. At pH 10, the Sm2CeMnO6/g-C3N4 composite achieved a 96.1% degradation rate for MB under visible light (λ > 365 nm) irradiation for 120 min. In the Sm2CeMnO6/g-C3N4 (20:80) composite, the two n-type semiconductors jointly formed an S-scheme heterojunction with customized valence- and conduction-band positions, demonstrating excellent performance in generating oxidative radicals. Wang et al. [58] developed a tubular carbon nitride (TCN)/titanium dioxide heterojunction photocatalytic system with S-scheme band alignment using a precursor structure reconstruction and stepwise calcination technique. As shown in Figure 3g, the synthesis process consisted of two critical stages: first, hollow tubular g-C3N4 substrate materials were prepared through a molecular precursor morphology control strategy [59], followed by the in situ growth of TiO2 nanocrystals on the TCN surface via a solvothermal method combined with calcination, ultimately achieving the precise construction of the heterointerface. The hierarchical pore topology structure endowed the material with a large specific surface area, significantly increasing the density of exposed active sites. Under visible light irradiation for 60 min, the tubular g-C3N4/TiO2 catalyst achieved a 96.6% degradation efficiency for MB. The performance improvement can be attributed to the enhanced light capture capability of the tubular structure and the improved charge transfer efficiency at the heterointerface. Barzegar et al. [60] fabricated S-scheme heterojunction g-C3N4/TiO2 composite materials via a straightforward sol–gel method. This composite material was employed in a solar parabolic trough reactor (PTC) to achieve the highly efficient degradation of MB. The g-C3N4/TiO2 heterojunction achieved a 94.9% degradation rate for MB under solar irradiation for 80 min. This outstanding performance was attributed to the S-scheme heterojunction: the conduction-band electrons of g-C3N4 and the valence-band holes of TiO2 were directionally migrated through the built-in electric field, which not only preserved the high redox capability but also effectively suppressed carrier recombination, thereby significantly enhancing photocatalytic activity. Riazati et al. [61] successfully developed a g-C3N4/CuO hetero-nanocomposite system using an in situ chemical synthesis strategy. The experimental results demonstrated that the composite catalyst loaded with 5 wt% CuO achieved a degradation efficiency of 93.0% for methylene blue after 180 min of visible light irradiation, with its apparent reaction kinetic constant being approximately six times higher than those of single-component g-C3N4 and CuO. Through a mechanism analysis, it was revealed that the enhanced photocatalytic performance of this system was ascribed to the built-in electric field formed by the band-bending effect at the heterointerface, which facilitated the formation of an S-scheme carrier transfer path. This transfer mode effectively inhibited the recombination of photogenerated carriers. In further studies, Ajami et al. [62] constructed a magnetically recyclable CoFe2O4/g-C3N4 heterostructure using a simple in situ chemical deposition method. This composite material exhibited optimal photocatalytic performance under visible light irradiation for 180 min, achieving 100.0% degradation of MB. The degradation pathway of the CoFe2O4/g-C3N4 nanocomposite followed an S-scheme mechanism. The heterojunction between cobalt ferrite and graphitic carbon nitride enabled improved light absorption, reduced bandgap energy, and enhanced charge carrier separation. Research on the degradation of MB using S-scheme heterojunction photocatalysts based on g-C3N4 has been conducted, with relevant studies reported in [27,29,48,55,56,57,58,60,61,62]. Detailed descriptions of these works are provided in Table 1.
Table 1. The photocatalytic degradation performance of g-C3N4-based S-scheme heterojunctions for MB.
Table 1. The photocatalytic degradation performance of g-C3N4-based S-scheme heterojunctions for MB.
PhotocatalystSynthesis MethodAmount of CatalystLight SourceAmount of MBTime
(min)
Efficiency
(%)
Ref.
GO/g-C3N4/TiO2Calcination
In situ crystallization
10 mg 350 W Xenon lamp,
λ ≥ 420 nm
100 mL, 10 mg/L24098.8[29]
Ag/AgI/g-C3N4Hydrothermal
Calcination
Photoreduction
/Simulated sunlight/5097.0[48]
g-C3N4/Co/ZnOUltrasound
Sol–gel
20 mg300 W Xenon lamp,
λ ≥ 420 nm
50 mL, 15 ppm12096.3[27]
CaSnO3/g-C3N4Solid-state route50 mg500 W Halogen lamp1 g/L12095.0[55]
PbTiO3/g-C3N4Ultrasound
Hydrothermal
50 mg100 W Halogen lamp 50 mL, 30 mg/L9099.4[56]
Sm2CeMnO6/g-C3N4Sol–gel150 mg9 W three LED lamp,
λ > 365 nm
100 mL, 8 mg/L12096.1[57]
TCN/TiO2Precursor reconstruction,
Hydrothermal,
Calcination
/Visible/6096.6[58]
g-C3N4/TiO2Sol–gel20 mg14.4 W/m LED lamp (SMD 5050 Flexible Strips, Ltd. China),
λ = 460 nm
10 mg/L8094.9[60]
g-C3N4/CuOIn situ synthesis200 mg150 W visible light source (Osram, Munich, Germany)50 mL, 2 mg/L18093.0[61]
CoFe2O4/g-C3N4In situ deposition200 mg/L150 W visible light source (Osram, Munich, Germany),
λ > 420 nm
50 mL, 2 mg/L180100.0[62]
Figure 3. (a) An SEM image of Ag/AgI40%/N-CT-10; a TEM image (b) and HRTEM image (c) of Ag/AgI40%/N-CT-10. (d) The ESR spectrum of Ag/AgI40%/N-CT-10. (e) The planar average electron density difference of the Ag/AgI 40%/N-CT-10 interlayer; (f) the charge distribution on Ag/AgI 40%/N-CT-10 (the cyan and yellow areas represent the depletion and accumulation of electrons, respectively) [48]. (g) A schematic diagram of the fabrication process of the TCN/TiO2 heterojunction structure [58].
Figure 3. (a) An SEM image of Ag/AgI40%/N-CT-10; a TEM image (b) and HRTEM image (c) of Ag/AgI40%/N-CT-10. (d) The ESR spectrum of Ag/AgI40%/N-CT-10. (e) The planar average electron density difference of the Ag/AgI 40%/N-CT-10 interlayer; (f) the charge distribution on Ag/AgI 40%/N-CT-10 (the cyan and yellow areas represent the depletion and accumulation of electrons, respectively) [48]. (g) A schematic diagram of the fabrication process of the TCN/TiO2 heterojunction structure [58].
Catalysts 15 00592 g003

4.1.2. Rhodamine B

Leelavathi et al. [27] successfully synthesized g-C3N4/Co/ZnO S-scheme heterojunction nanocomposites via a simple ultrasonic-assisted sol–gel method. The g-C3N4/Co/ZnO nanocomposite achieved a photocatalytic degradation efficiency of 75.1% for RhB under visible light irradiation for 120 min, and the degradation efficiency increased to 91.6% under solar light irradiation for 120 min, as shown in Table 2. In-depth analysis revealed that the local surface plasmon resonance (LSPR) effect of Co nanoparticles significantly enhanced the local electric field and improved the photocatalyst’s responsiveness to sunlight. Sun et al. [48] developed Ag/AgI/g-C3N4 S-scheme heterojunction photocatalysts by integrating hydrothermal synthesis, calcination, and photoreduction techniques. The Ag/AgI40%/N-CT-10 photocatalyst exhibited a degradation efficiency of 95.6% for RhB under simulated solar light irradiation for 50 min. The synergistic enhancement of the nitrogen-vacancy effect and the local surface plasmon resonance effect in Ag/AgI40%/N-CT-10 resulted in superior photocatalytic degradation performance. Alsalme et al. [63] fabricated Zr(HPO4)2/g-C3N4 S-scheme heterojunctions using an ultrasonic chemical coupling method. The Zr(HPO4)2/g-C3N4 photocatalyst achieved a degradation rate of 98.0% for RhB under simulated natural solar light irradiation at 1000 W power for 3 h. This excellent performance was attributed to the S-scheme charge transfer mechanism. Chopan et al. [64] successfully prepared an S-scheme photocatalytic system of g-C3N4/PPy/ZnO through a multi-step process involving hydrothermal synthesis, calcination, and polymerization. The g-C3N4/PPy/ZnO photocatalyst achieved a degradation rate of 99.0% for RhB under visible light irradiation for 60 min. The outstanding photocatalytic activity of g-C3N4/PPy/ZnO was ascribed to the oxygen vacancies (OVs) on ZnO, promoting the effective separation of photogenerated carriers and enhancing their redox ability through an S-scheme charge transfer pathway during heterojunction formation. Shoaib et al. [65] successfully fabricated a double S-scheme CdS/TiO2/g-C3N4 heterojunction via a hydrothermal method. This material achieved a degradation efficiency of 99.4% for RhB under visible light irradiation for 180 min, significantly surpassing binary and single-component catalysts. The double S-scheme structure enabled directional charge transfer between CdS and g-C3N4 through TiO2 as a charge bridge, preserving the high-conduction-band electrons of CdS (–0.37 eV) and the strong reduction capability of g-C3N4, while the high-valence-band holes of TiO2 (+2.93 eV) enhanced the oxidation reaction efficiency.
Patra et al. [56] synthesized a PbTiO3/g-C3N4 S-scheme heterojunction using an ultrasound-assisted hydrothermal method. The PbTiO3/g-C3N4 photocatalyst achieved a degradation efficiency of 99.8% for RhB under simulated solar radiation for 60 min. Its superior performance was attributed to the S-scheme charge transfer mechanism. Barzegar et al. [60] prepared an S-scheme heterojunction g-C3N4/TiO2 composite material via an ultrasonic method and applied it for continuous RhB degradation in a solar parabolic trough reactor. The g-C3N4/TiO2 composite material achieved a degradation efficiency of 93.1% for RhB under solar irradiation for 80 min. This outstanding performance was ascribed to the S-scheme heterojunction: the conduction-band electrons of g-C3N4 and the valence-band holes of TiO2 were directionally recombined at the heterojunction interface, retaining the strong reduction capability of g-C3N4 and the high oxidation ability of TiO2. Meanwhile, the solar parabolic trough reactor significantly improved light utilization efficiency by focusing solar energy. Alsalme et al. [66] synthesized an S-scheme heterojunction Ag2CrO4/g-C3N4 composite material through an acoustic chemical method. The 20 wt% Ag2CrO4/g-C3N4 composite material degraded 96.0% of RhB under sunlight irradiation for 2 h, outperforming pure g-C3N4 (60.0%) and Ag2CrO4 (33.0%). Radical-trapping experiments revealed that superoxide radicals and holes were the primary active species. The S-scheme heterojunction preserved the high-valence-band holes of Ag2CrO4 (+2.21 eV) and the high-conduction-band electrons of g-C3N4 (–1.3 eV) through directional charge recombination at the interface, synergistically enhancing the separation efficiency of photogenerated carriers and redox capability. Alsalme et al. [67] successfully synthesized S-scheme AgI/g-C3N4 heterojunctions via a sonochemical method. The 10 wt% AgI/g-C3N4 composite material achieved degradation efficiencies of 96.0% for RhB and 86.0% for tetracycline under sunlight irradiation for 2 h, significantly outperforming pure g-C3N4 (12.0%) and AgI (40.0%). This enhanced performance was attributed to the S-scheme heterojunction, which facilitated directional charge recombination at the interface, preserving the high-valence-band holes of AgI (+2.28 eV) and the high-conduction-band electrons of g-C3N4 (–1.2 eV). Rajendran et al. [68] developed S-scheme g-C3N4/TiO2/CuO ternary heterojunctions using a hydrothermal method. The g-C3N4/TiO2-CuO composite material demonstrated a degradation efficiency of 90.3% for RhB under simulated sunlight irradiation for 120 min, surpassing that of pure g-C3N4 (31.8%). The superior performance was ascribed to the S-scheme heterojunction: the conduction-band electrons of g-C3N4 and the valence-band holes of TiO2/CuO were directionally recombined at the interface, retaining the strong reduction ability of g-C3N4 and the high oxidation ability of CuO/TiO2 while simultaneously enhancing the light absorption range and charge separation efficiency. Leelavathi et al. [27] fabricated S-scheme g-C3N4/Co/ZnO heterojunctions through an ultrasonic-assisted sol–gel method. This material exhibited degradation efficiencies of 96.3%, 74.5%, and 75.1% for MB, crystal violet, and RhB, respectively, under visible light irradiation for 80 min. Under sunlight irradiation, the degradation efficiency of RhB increased to 91.6%. This outstanding performance was attributed to the S-scheme heterojunction: the surface plasmon resonance effect of Co nanoparticles enhanced light absorption, and the conduction-band electrons of g-C3N4 and the valence-band holes of ZnO were directionally recombined at the interface, preserving the strong reduction ability of g-C3N4 and the high oxidation ability of ZnO while simultaneously improving the charge separation efficiency. Alsulmi et al. [69] successfully fabricated S-scheme CeO2/g-C3N4 heterojunctions using an ultrasonic-assisted sol–gel method. The 15 wt% CeO2/g-C3N4 composite material achieved a degradation efficiency of 98.9% for RhB under natural light irradiation for 2 h, with a rate constant of 0.0312 min⁻1, which was 16.4 times higher than that of pure g-C3N4. This outstanding performance was attributed to the S-scheme heterojunction: the high-valence-band holes of CeO2 (+2.37 eV) and the high-conduction-band electrons of g-C3N4 (–1.2 eV) were directionally recombined at the interface, preserving the strong oxidation–reduction capabilities of both materials. Additionally, the oxygen vacancies in CeO2 enhanced light absorption and charge separation efficiency. Alsalme et al. [63] synthesized S-scheme Zr(HPO4)2/g-C3N4 heterojunctions via an ultrasonic chemical method. The 5 wt% Zr(HPO4)2/g-C3N4 composite material achieved a degradation efficiency of 98% for RhB under sunlight irradiation for 3 h, with a rate constant of 0.048 min⁻1. This excellent performance was ascribed to the S-scheme heterojunction: the high-valence-band holes of Zr(HPO4)2 (+3.0 eV) and the high-conduction-band electrons of g-C3N4 (–1.25 eV) were directionally recombined at the interface, preserving the strong oxidation–reduction capabilities of both materials. Furthermore, ultrasonic treatment promoted the uniform dispersion of nanoparticles, enhancing the light absorption and charge separation efficiency. Alsalme et al. [70] compared the preparation of S-scheme SnS2/g-C3N4 heterojunctions using ultrasonic-assisted and physical mixing methods. The 15 wt% SnS2/g-C3N4 composite material prepared using the ultrasonic method achieved a degradation efficiency of 98.0% for RhB under sunlight irradiation for 2 h, significantly surpassing the physical mixing method (55.0%). The S-scheme structure facilitated the directional recombination of the high-valence-band holes of SnS2 (+2.15 eV) and the high-conduction-band electrons of g-C3N4 (–1.2 eV) at the interface, preserving the strong oxidation–reduction capabilities of both materials. Ultrasonic treatment also promoted the uniform dispersion of SnS2, suppressing charge recombination and enhancing the light absorption efficiency. Alsulmi et al. [71] developed S-scheme Ag2CO3/g-C3N4 heterojunctions via an ultrasonic chemical method. The 5 wt% Ag2CO3/g-C3N4 composite material achieved a degradation efficiency of 95.0% for RhB under natural light irradiation for 2 h, with a rate constant of 0.0141 min⁻1. This superior performance was attributed to the S-scheme heterojunction: the high-valence-band holes of Ag2CO3 (+2.55 eV) and the high-conduction-band electrons of g-C3N4 (–1.13 eV) were directionally recombined at the interface, preserving the strong oxidation–reduction capabilities of both materials. Ultrasonic treatment further promoted the uniform dispersion of nanoparticles, enhancing the light absorption and charge separation efficiency. Huang et al. [72] successfully fabricated S-scheme HKUST-1 (copper-based MOFs)/g-C3N4 heterojunctions via an ultrasonic-assisted method for the degradation of RhB driven by the synergistic effect of visible light and ultrasound irradiation. The 30% HC (HKUST-1/g-C3N4) composite achieved a degradation efficiency of 94.4% for RhB after 120 min of visible light irradiation, with a rate constant of 0.0224 min−1. This outstanding performance was attributed to the S-scheme heterojunction: the high-valence-band holes of HKUST-1 (+2.43 eV) and the high-conduction-band electrons of g-C3N4 (–1.21 eV) were directionally recombined at the interface, preserving the strong oxidation–reduction capabilities of both materials. Additionally, the piezoelectric polarization effect induced by ultrasound further enhanced the charge separation efficiency. Basely et al. [73] synthesized S-scheme Bi2S3/g-C3N4 heterojunctions via an ultrasonic chemical method. The Bi2S3/g-C3N4 composite material achieved a degradation efficiency of 90.0% for RhB under natural light irradiation for 2 h, with a rate constant of 0.0073 min⁻1. Its excellent performance was ascribed to the S-scheme heterojunction: the high-valence-band holes of Bi2S3 (+2.04 eV) and the high-conduction-band electrons of g-C3N4 (–1.34 eV) were directionally recombined at the interface, preserving the strong oxidation–reduction capabilities of both materials. Furthermore, ultrasonic treatment promoted the uniform dispersion of nanoparticles, enhancing light absorption and charge separation efficiency. Alsulmi et al. [67] developed S-scheme AgI/g-C3N4 heterojunctions via an ultrasonic chemical method. The 10 wt% AgI/g-C3N4 composite material achieved degradation efficiencies of 96.0% for RhB and 86.0% for tetracycline under natural light irradiation for 2 h. Its superior performance was attributed to the S-scheme heterojunction: the high-valence-band holes of AgI (+2.28 eV) and the high-conduction-band electrons of g-C3N4 (–1.2 eV) were directionally recombined at the interface, preserving the strong oxidation–reduction capabilities of both materials. Ultrasonic treatment also promoted the uniform dispersion of nanoparticles, enhancing light absorption and charge separation efficiency. Elavarasan et al. [74] successfully synthesized a double S-scheme g-C3N4/rGO/ZnO-Ag heterojunction via a hydrothermal method. The HRTEM image in Figure 4a reveals that g-C3N4 exhibits a sheet-like structure, rGO displays a layered morphology, and ZnO/Ag presents as spherical agglomerates [75,76]. Additionally, some dark spherical particles can be observed, which are uniformly dispersed within the active ZnO-Ag composite material. The photocatalytic degradation performance of the g-C3N4/rGO/ZnO-Ag composite material was evaluated under visible light irradiation for 100 min, with RhB and MB dyes in a mixed aqueous solution serving as the target pollutants. The results demonstrated that the degradation efficiencies of the composite material for RhB and MB were 83.4% and 90.0%, respectively. Figure 4b illustrates the degradation of the RhB/MB mixed dye system by the composite material, with the dye concentration changes monitored at 552 nm (characteristic absorption peak of RhB) and 664 nm (characteristic absorption peak of MB) using spectrophotometry. Figure 4c elucidates the S-scheme charge transfer mechanism. Metal silver nanoparticles (AgNPs) significantly enhance the light energy capture efficiency in the visible light spectrum through the surface plasmon resonance effect. This effect not only induces the formation of a high-intensity localized electromagnetic field but also promotes the excitation kinetics of photogenerated e⁻-h⁺ pairs in semiconductor materials [77]. Upon light irradiation, g-C3N4 is excited by photons to generate charge carriers. Notably, the plasmonic resonance characteristics of AgNPs optimize the visible light absorption performance, thereby facilitating the efficient transfer of photogenerated electrons to the conduction band of ZnO and effectively suppressing carrier recombination. The spatial charge separation mechanism at the heterojunction interface regulates the dynamic balance of photogenerated e⁻-h⁺ pairs, significantly reducing their recombination probability and enhancing photocatalytic reaction efficiency. Moreover, the introduction of reduced graphene oxide as an electron transport medium creates a rapid charge transfer channel, optimizing the carrier separation pathway and synergistically enhancing photocatalytic activity.
Sharma et al. [78] successfully fabricated S-scheme α-Fe2O3/g-C3N4/SiO2 heterojunctions via hydrothermal and calcination methods for the visible-light-assisted photo-Fenton degradation of RhB. Under the conditions of pH = 3, a 0.6 g/L catalyst loading, and 7 × 10⁻4 M H2O2, the α-Fe2O3/g-C3N4/SiO2 photocatalyst achieved a degradation efficiency of 97% for RhB after 120 min of visible light irradiation. The enhanced photocatalytic performance of the Fe2O3/g-C3N4/SiO2 S-scheme heterojunction is attributed to the efficient interface charge separation and band-bending structure between g-C3N4 and α-Fe2O3, facilitated by the supporting role of SiO2, which significantly improves the separation efficiency of photogenerated carriers. Bui et al. [79] developed an S-scheme heterojunction composed of oxygen-vacancy-deficient SnO2 nanoparticles and g-C3N4 via calcination and hydrothermal methods. Activated by potassium peroxymonosulfate (PMS) in solution, the S-scheme structure of SnO2/g-C3N4 achieved a degradation efficiency of 99.8% for RhB after 180 min of visible light irradiation. The S-scheme heterojunction promotes the directional recombination of the high-valence-band holes of SnO2 (+3.78 eV) and the high-conduction-band electrons of g-C3N4 (–1.13 eV) through oxygen vacancies, preserving their strong redox capabilities. Meanwhile, the sulfate radicals (SO4⁻) and •OH generated by PMS activation synergistically enhance the degradation efficiency. Mishra et al. [80] prepared S-scheme g-C3N4/NiFe2O4 heterojunctions via sol–gel combustion combined with calcination for the visible-light-assisted persulfate (PS) degradation of RhB and tetracycline. The g-C3N4/NiFe2O4 composite achieved degradation efficiencies of 98.6% and 84.3% for RhB and tetracycline, respectively, after 60 min of visible light irradiation, with rate constants of 0.027 min⁻1 and 0.018 min⁻1. The S-scheme heterojunction preserves strong redox capabilities by promoting the directional recombination of the conduction-band electrons of g-C3N4 (–1.79 eV) and the valence-band holes of NiFe2O4 (+2.22 eV). Additionally, the Fe2⁺/Fe3⁺ redox cycle facilitates PS activation, generating SO4⁻ and •OH, thereby enhancing the overall catalytic efficiency. Liang et al. [81] successfully fabricated an S-scheme heterostructure of potassium-doped carbon nitride (KCN) and ZnO in a KCl/LiCl molten salt system via a high-temperature melting method. Transmission electron microscopy (TEM) analysis (Figure 4d) revealed the disordered assembly morphology of nanorods in the composite, confirming that the molten-salt-induced phase transformation facilitated a topological transition from two-dimensional sheets to one-dimensional nanorods. High-resolution lattice imaging (Figure 4e) revealed a crystal plane spacing of 0.284 nm, which perfectly matched the (100) crystal plane of ZnO [82], while periodic stripes with a spacing of 0.312 nm corresponded to the characteristic parameters of the (100) crystal plane of KCN [83]. X-ray diffraction (XRD) patterns further confirmed the presence of the characteristic diffraction peaks of polytriazine imide (PTI), polyheptazine imide (PHI), and ZnO, verifying the successful synthesis of the ternary composite material [84]. This structural configuration provided a basis for the spatial separation of photogenerated carriers, significantly enhancing the catalytic efficiency of the composite system. Nitrogen adsorption–desorption isotherms (Figure 4f) demonstrated that all samples exhibited type IV isotherms accompanied by H3-type hysteresis loops, indicating that the materials possessed typical mesoporous characteristics [85]. Quantitative analysis (Figure 4g) indicated that the specific surface area and pore volume of KCN/ZnO were notably increased compared to those of the individual components. This hierarchical pore structure not only enhanced light harvesting but also significantly increased the density of active surface sites. Additionally, the average pore diameter of the composite material was intermediate between those of the original components, forming a pore size gradient conducive to reactant diffusion. Photocatalytic evaluation (Figure 4h) showed that KCN/ZnO achieved a degradation efficiency of 92.0% for Rhodamine B under full-spectrum irradiation within 120 min, demonstrating a significant synergistic effect compared to the individual effects of KCN (55.1%) and ZnO (37.2%). Kinetic analysis (Figure 4i) revealed that the degradation process followed a pseudo-first-order reaction model, with the apparent rate constant (k value) of the composite material being 5–8 times higher than those of the individual components, directly validating the mechanism of enhanced catalytic activity [86]. The S-scheme carrier transport channel constructed at the heterointerface not only significantly improved the charge separation efficiency but also retained a high redox potential through band alignment. Sun et al. [87] successfully fabricated a three-dimensional porous sulfur-doped graphitic carbon nitride composite TiO2/SiO2/PAN aerogel (S-gTAHP) heterostructure via electrospinning and hydrothermal–freeze-drying techniques. Morphological characterization (Figure 5a) revealed that, after 15 h of hydrothermal treatment (S-gTAHP-15h), the aerogel exhibited a highly ordered pore distribution. This optimized multi-level pore framework not only increased the specific surface area but also provided favorable conditions for the exposure of active sites. Microstructural analysis indicated that TiO2/PAN short fibers formed a three-dimensional skeleton, while sulfur-doped carbon nitride nanosheets were uniformly embedded in the matrix through SiO2 crosslinking agents. The UV–Vis absorption spectrum (Figure 5b) demonstrated that TAP-15h (TiO2/SiO2/PAN aerogel) had a distinct absorption edge at 390 nm, corresponding to a bandgap value of 3.16 eV (Figure 5c). In contrast, the absorption edge of S-gAP-15h (sulfur-doped g-C3N4/SiO2 aerogel) redshifted to 460 nm, with a narrowed band gap of 2.71 eV. Notably, the ternary composite system S-gTAHP exhibited significantly enhanced absorption characteristics in the visible light region, attributed to the regulation of the band structure via the synergistic effect of sulfur doping and the porous network on light harvesting [88]. As the hydrothermal time increased from 9 to 15 h, the absorption edge of the material continuously redshifted, confirming that the gradual improvement in the three-dimensional porous structure enhanced light energy utilization through multiple light scattering mechanisms. The carrier dynamics were investigated using photoluminescence spectra (Figure 5d) and electrochemical testing systems [89]. S-gTAHP-15h exhibited the lowest fluorescence intensity in the PL spectrum, indicating that the close contact at the heterojunction interface effectively suppressed electron–hole recombination. The transient photocurrent response (Figure 5e) demonstrated that this material had excellent photocurrent response reproducibility, with a photocurrent density value significantly superior to that of the comparison samples. Electrochemical impedance spectroscopy (Figure 5f) further confirmed that S-gTAHP-15h had the smallest capacitive arc diameter, indicating the lowest interfacial charge transfer resistance. These characterization results were consistent with the photocatalytic performance data: under simulated solar light irradiation, this material exhibited excellent degradation efficiency for methylene blue (99.4% degradation rate in 15 min), Rhodamine B (96.1% in 30 min), and tetracycline (84.2% in 40 min). The performance improvement mechanism can be summarized as follows: (1) the three-dimensional multi-level pore structure enhances light energy capture by increasing the density of active sites and extending the light propagation pathway; (2) the efficient carrier migration channels established by the S-scheme heterostructure promote charge separation; (3) the sulfur doping strategy effectively expands the light response range of the material. Pitcheri et al. [90] successfully fabricated an S-scheme heterojunction β-Cu2V2O₇/Ni/Pg-C3N4 composite photocatalytic system by employing a simple co-precipitation process to directionally load β-Cu2V2O₇ nanoparticles onto porous graphitic carbon nitride (Pg-C3N4) nanosheets, with the interfacial modification effect of nickel nanoparticles. Photocatalytic performance tests revealed that this composite material achieved a removal efficiency of 98.6% for RhB after 60 min of sunlight exposure, with a reaction rate constant (0.039 min⁻1) that was 4.5 times higher than that of pure β-Cu2V2O₇. The performance enhancement mechanism can be attributed to the following synergistic effects: (1) the built-in electric field formed at the interface between Pg-C3N4 nanosheets and β-Cu2V2O₇ in the S-scheme heterojunction effectively regulates the carrier migration path; (2) the surface plasmon resonance effect of nickel nanoparticles enhances the excitation and separation of photogenerated electron–hole pairs through a localized electromagnetic field; (3) unique heterointerface engineering significantly prolongs the lifetime of photogenerated charges. Madima et al. [91] developed a direct S-scheme TiO2/g-C3N4 heterojunction by simultaneously calcining TiO2 precursors and g-C3N4. TiO2 was synthesized using guava leaf extract as a reducing agent via a green synthesis method, while g-C3N4 was prepared via the thermal decomposition of melamine. This composite material degraded 96.0% of RhB within 120 min and 95.0% of MB within 150 min under simulated sunlight irradiation. The enhanced photocatalytic activity was attributed to the visible light capture characteristics and the S-scheme heterojunction system formed between the two catalysts, which promoted interfacial charge separation efficiency and prolonged the charge carrier lifetime. Khamesan et al. [92] prepared a 2D/2D S-scheme g-C3N4/ZnCr-LDH heterojunction by combining hydrothermal and reflux methods for the degradation of RhB and the in situ production of H2O2 under xenon lamp irradiation. The g-C3N4/ZnCr-LDH composite achieved a degradation efficiency of 99.8% for RhB and produced 4.75 mmol/L of H2O2 after 90 min of simulated sunlight exposure. Comprehensive experimental results, bandgap energy analysis, and scavenging tests confirmed the presence of an S-scheme heterojunction in the synthesized g-C3N4/ZnCr-LDH composite material. The proposed S-scheme mechanism accelerates the separation, migration, and utilization of photogenerated charge carriers, thereby enhancing its photocatalytic activity. Xie et al. [93] successfully fabricated an S-scheme g-C3N4/Bi/BiVO4 composite photocatalytic system through a synergistic strategy combining chemical deposition and in situ reduction. High-resolution transmission electron microscopy analysis (Figure 5g) revealed that monoclinic-phase BiVO4 nanosheets (m-BiVO4) formed a heterojunction with g-C3N4 nanosheets via the (001) crystal plane, with their surfaces being uniformly covered by amorphous Bi nanoparticles. The (121) crystal plane diffraction fringes of m-BiVO4 were resolved. Energy-dispersive X-ray spectroscopy (EDS) elemental mapping (Figure 5h) further confirmed the homogeneous distribution of C, N, Bi, V, and O elements within the composite system, validating the effective integration of material interfaces. Under visible light irradiation for 70 min, the composite catalyst achieved a degradation efficiency of 99.0% for Rhodamine B, with a reaction rate constant of 0.067 min⁻1, demonstrating significant kinetic advantages. The performance enhancement mechanism can be attributed to (1) the establishment of a directional charge transfer channel by the S-scheme heterojunction through band alignment, preserving the high redox potentials of the charge carriers; (2) the role of Bi nanoparticles as electron relay stations, enhancing interfacial charge vector migration. Research on the degradation of RhB using S-scheme heterojunction photocatalysts based on g-C3N4 has been reported in several studies, including [27,48,56,60,63,64,65,66,67,68,69,70,71,72,73,74,78,79,80,81,90,91,92,93]. Detailed information regarding these works is summarized in Table 2.
Table 2. The photocatalytic degradation performance of g-C3N4-based S-scheme heterojunctions for RhB.
Table 2. The photocatalytic degradation performance of g-C3N4-based S-scheme heterojunctions for RhB.
PhotocatalystSynthesis MethodLight SourceAmount of RhBAmount of CatalystTime
(min)
Efficiency
(%)
Ref.
g-C3N4/Co/ZnOUltrasound, sol–gel300 W Xenon lamp,
λ > 420 nm
50 mL, 15 ppm20 mg 12075.1[27]
Ag/AgI/g-C3N4Hydrothermal, calcination,
photoreduction
Simulated sunlight//5095.6[48]
Zr(HPO4)2/g-C3N4Ultrasonic chemical coupling1000 W sunlight100 mL, 2×10−5 M 10 mg18098.0[63]
g-C3N4/PPy/ZnOHydrothermal, calcination,
polymerization
125 W LED lamp,
λ ≥ 420 nm
100 mL, 50 mg/L100 mg6099.0[64]
CdS/TiO2/g-C3N4Hydrothermal50 W LED lamp50 mL, 25 mg/L10 mg18099.4[65]
PbTiO3/g-C3N4Ultrasound, hydrothermal100 W halogen lamp50 mL, 10 mg/L50 mg6099.8[56]
g-C3N4/TiO2Ultrasound14.4 W/m LED lamp (SMD 5050 Flexible Strips, Ltd. China),
λ = 460 nm
15 mg/L20 mg8093.1[60]
Ag2CrO4/g-C3N4Ultrasound300 W solar simulator100 mL, 2 × 10−5 M50 mg12096.0[66]
AgI/g-C3N4Ultrasound300 W solar simulator100 mL, 2 × 10−5 M100 mg12096.0[67]
g-C3N4/TiO2/CuOHydrothermal500 W Xenon lamp100 mL, 30 ppm50 mg12090.3[68]
CeO2/g-C3N4Ultrasound, sol–gel110 K to 90 K lumen, natural solar100 mL, 2 × 10−5 M20 mg12098.9[69]
SnS2/g-C3N4Ultrasound1000 W natural solar100 mL, 2 × 10−5 M50 mg12098.0[70]
Ag2CO3/g-C3N4Ultrasound450 W natural sunlight100 mL, 10 mg/L100 mg12095.0[71]
HKUST-1/g-C3N4UltrasoundUltraviolet–visible light source50 mL, 10 mg/L 50 mg12094.4[72]
Bi2S3/g-C3N4Ultrasound500 W natural sunlight//12090.0[73]
g-C3N4/rGO/ZnO-AgHydrothermalVisible100 mL, 40 ppm60 mg10083.4[74]
α-Fe2O3/g-C3N4/SiO2Hydrothermal, calcination100 W LED lamp,
λ = 420 nm
100 mL, 10 ppm60 mg12097.0[78]
SnO2/g-C3N4Calcination, hydrothermalVisible light source,
λ > 420 nm
20 mg/L0.33 g/L15099.8[79]
g-C3N4/NiFe2O4Sol–gel combustion, calcinationSunlight50 mL, 20 mg/L20 mg6098.6[80]
K-g-C3N4/ZnOHigh-temperature meltingVisible//12092.0[81]
β-Cu2V2O₇/Ni/Pg-C3N4Co-precipitationSunlight60 mL, 25 ppm 20 mg 6098.6[90]
TiO2/g-C3N4Calcination300 W Xenon lamp (Asahi HAL-320)100 mL, 10 mg/L50 mg 12096.0[91]
g-C3N4/ZnCr-LDHHydrothermal, refluxLED lamp100 mL, 5 ppm100 mg9099.8[92]
g-C3N4/Bi/BiVO4Deposition, in situ reduction350 W Xenon lamp (Changzhou Siyu, China),
λ > 420 nm
50 mL, 10 mg/L 50 mg7099.0%[93]
Figure 4. (a) An HRTEM image of the GCRZA S-scheme heterojunction photocatalyst; (b) the photocatalytic activity of the GCRZA nanocomposite toward RhB + MB mixed dyes; (c) a schematic diagram of the photocatalytic degradation process of the GCRZA S-scheme heterojunction photocatalyst [74]; (d) TEM images of the KCN/ZnO nanocomposite; (e) XRD patterns, (f) N2 adsorption–desorption isotherms, and (g) corresponding pore size distributions of the KCN/ZnO nanocomposite, KCN, and ZnO; the photocatalytic degradation performance of the KCN/ZnO nanocomposite, KCN, and ZnO for RhB and (h) the corresponding photocatalytic degradation kinetics (i) [81].
Figure 4. (a) An HRTEM image of the GCRZA S-scheme heterojunction photocatalyst; (b) the photocatalytic activity of the GCRZA nanocomposite toward RhB + MB mixed dyes; (c) a schematic diagram of the photocatalytic degradation process of the GCRZA S-scheme heterojunction photocatalyst [74]; (d) TEM images of the KCN/ZnO nanocomposite; (e) XRD patterns, (f) N2 adsorption–desorption isotherms, and (g) corresponding pore size distributions of the KCN/ZnO nanocomposite, KCN, and ZnO; the photocatalytic degradation performance of the KCN/ZnO nanocomposite, KCN, and ZnO for RhB and (h) the corresponding photocatalytic degradation kinetics (i) [81].
Catalysts 15 00592 g004

4.1.3. Other Cationic Dyes

Leelavathi et al. [27] successfully fabricated a ternary S-scheme heterojunction composite system of g-C3N4/Co/ZnO via an ultrasonic-assisted sol–gel method. The experimental results demonstrated that the catalyst achieved a removal efficiency of 74.5% for crystal violet under visible light irradiation for 120 min, as shown in Table 3. The study revealed that the plasmonic effect of cobalt nanoparticles effectively promoted the directional migration of carriers and significantly extended the material’s light response range. Sun et al. [48] developed an Ag/AgI/g-C3N4 S-scheme heterojunction photocatalyst by integrating hydrothermal synthesis, calcination, and photoreduction techniques. Under simulated sunlight irradiation for 60 min, Ag/AgI40%/N-CT-10 exhibited a degradation efficiency of 100.0% for crystal violet. The enhanced photocatalytic performance was attributed to the synergistic effects of nitrogen vacancies and localized surface plasmon resonance in Ag/AgI40%/N-CT-10. Mallah et al. [94] synthesized Ca@TiO2@g-C3N4 nanocomposites through a thermal polymerization reaction of non-noble-metal components followed by ultrasonic treatment. Under visible light irradiation for 60 min, the degradation of malachite green reached 99.9%. The high activity of this ternary photocatalyst can be ascribed to the formation of an S-scheme heterojunction at the TiO2@g-C3N4 interface, the generation of defects such as Ti3⁺ and oxygen vacancies (Oₛ) due to the interaction between Ca and TiO2, and the increased specific surface area of the photocatalyst. Madonna et al. [95] constructed a C-CeO2/g-C3N4 stepwise heterojunction via a hydrothermal method. Under sunlight irradiation for 150 min, the C-CeO2/g-C3N4 nanocomposite degraded 91.9% of malachite green. Comprehensive characterization confirmed that the enhanced photocatalytic activity of this material was primarily attributable to the effective establishment of an S-scheme charge transfer mechanism, which provided a unique carrier separation pathway and significantly suppressed electron–hole recombination.
In summary, g-C3N4-based S-scheme heterojunctions have exhibited remarkable performance and hold broad application prospects in the degradation of cationic dyes. By employing various advanced preparation techniques (hydrothermal synthesis, calcination, sol–gel processing, electrospinning, etc.), researchers have successfully developed a range of highly efficient S-scheme heterojunction photocatalytic systems. These systems have achieved significant improvements in the separation efficiency of photogenerated electron–hole pairs, expanded the spectral response range, and enhanced the efficiency of reactive oxygen species generation through the optimization of band structure design and the regulation of carrier migration mechanisms. In degradation experiments involving methylene blue, Rhodamine B, and other cationic dyes, g-C3N4-based S-scheme heterojunction composite materials have consistently demonstrated superior photocatalytic performance. Furthermore, studies have elucidated the substantial enhancement effects of various synergistic phenomena (oxygen vacancy engineering, sulfur doping, surface plasmon resonance, etc.) on photocatalytic activity. These findings provide novel insights for designing efficient and stable photocatalytic materials.
Figure 5. (a) An SEM image of the S-gTAHP-15 h aerogel photocatalyst; (b) the UV–Vis absorbance spectra of the photocatalyst and (c) Tauc’s plots of TAP-15 h and S-gAP-15 h photocatalysts; (d) the PL spectra, (e) transient I-t curves, and (f) EIS curves and corresponding equivalent circuit diagrams (illustrations) of different types of aerogel-based photocatalysts [87]. (g) An HRTEM image and (h) the corresponding EDS elemental maps of the g-C3N4/Bi/BiVO4 photocatalyst [93]. (i) The rate constants for the photocatalytic degradation of MO by different photocatalysts [96].
Figure 5. (a) An SEM image of the S-gTAHP-15 h aerogel photocatalyst; (b) the UV–Vis absorbance spectra of the photocatalyst and (c) Tauc’s plots of TAP-15 h and S-gAP-15 h photocatalysts; (d) the PL spectra, (e) transient I-t curves, and (f) EIS curves and corresponding equivalent circuit diagrams (illustrations) of different types of aerogel-based photocatalysts [87]. (g) An HRTEM image and (h) the corresponding EDS elemental maps of the g-C3N4/Bi/BiVO4 photocatalyst [93]. (i) The rate constants for the photocatalytic degradation of MO by different photocatalysts [96].
Catalysts 15 00592 g005

4.2. The Application of S-Scheme Heterojunction g-C3N4 Structures in the Degradation of Anionic Dyes

In recent years, g-C3N4-based S-scheme heterojunctions have exhibited remarkable potential in the degradation of anionic dyes, attributed to their distinctive charge separation mechanism and band structure regulation characteristics. In contrast to traditional photocatalytic systems, S-scheme heterojunctions enable the highly efficient spatial separation of photogenerated electron–hole pairs via the built-in electric field and band-bending effect at the interface while preserving the robust oxidation–reduction capabilities of each component. This significantly enhances both photocatalytic activity and the spectral response range. In studies focused on the degradation of anionic dyes such as methyl orange (MO), Congo red (CR), indigo carmine (IC), chrome black T (EBT), alizarin red S (ARS), and erythrosine (ER), researchers have successfully developed composite systems with superior photocatalytic efficiency through strategies including multi-component composites, defect engineering, and three-dimensional structural design. This section systematically reviews the latest advancements in S-scheme heterojunction g-C3N4 systems for the degradation of anionic dyes, emphasizing the optimization of band structures, interfacial charge transfer mechanisms, and multi-component synergistic effects, thereby providing valuable technical insights for the rational design of environmental remediation materials.

4.2.1. Methyl Orange

Rana et al. [96] innovatively integrated a cork substrate with the ternary components of g-C3N4/ZnO/TiO2 through a co-precipitation method combined with a physical composite process, thereby constructing a multiphase photocatalytic system with S-scheme heterojunction characteristics. Photocatalytic performance tests demonstrated that the g-C3N4/ZnO/TiO2/cork composite material achieved a removal efficiency of 98.3% for methyl orange within 60 min of visible light irradiation, as shown in Table 3. Kinetic analysis (Figure 5i) revealed a significant cascade effect in the photocatalytic activity of each component: g-C3N4/ZnO/TiO2/cork composite system (0.0524 min⁻1) > g-C3N4/ZnO/TiO2 (0.0258 min⁻1) > g-C3N4/ZnO (0.0155 min⁻1) > g-C3N4 (0.0121 min⁻1) > ZnO (0.0107 min⁻1) > TiO2 (0.00925 min⁻1); this fully verified the synergistic enhancement effect of the cork substrate and the multi-component heterojunction. Systematic studies of radical scavenging experiments revealed that when Na2-EDTA, isopropanol (IPA), and benzoquinone (BQ) were added, the degradation efficiency of MO decreased to 61.3%, 24.5%, and 27.9%, respectively, under the same conditions (Figure 6a). This confirmed the dominant role of h⁺, •O2⁻, and •OH radicals in the photocatalytic process. Mechanism studies indicated that the special band structure of the double S-scheme heterojunction not only effectively suppressed the recombination of photogenerated electron–hole pairs but also promoted the separation of photogenerated carriers via a unique carrier migration pathway. The multi-component synergy effect further broadened the spectral response range. This interface engineering strategy provides a new design concept for the development of highly efficient and stable composite photocatalysts. Gou et al. [97] successfully fabricated a CoTiO3/g-C3N4 heterojunction photocatalytic system by employing an in situ calcination strategy using ZIF-67@TiO2 and melamine precursors (synthesis process shown in Figure 6b). The CoTiO3/g-C3N4-2 sample achieved a degradation efficiency of 99.7% for methyl orange after 4 h of visible light irradiation (λ > 420 nm). Its superior performance can be attributed to (1) the directional carrier migration mechanism of the S-scheme heterojunction, which retains strong redox capability while expanding the solar spectral response range; (2) the precise matching of the band structure at the heterointerface, which optimizes the separation efficiency of photogenerated charges.
Onwudiwe et al. [98] constructed a ternary S-scheme heterojunction of g-C3N4/Bi2S3/CuS via a solvothermal method. This system significantly enhanced carrier separation and suppressed recombination through band engineering. When the loading amount of CuS was optimized to 20%, the removal of methyl orange by the composite material increased to 98.0% within 60 min of visible light irradiation. Experimental characterization and theoretical calculations indicated that the built-in electric field within the heterojunction and the multi-component energy level had a synergistic effect, jointly promoting the effective separation of photogenerated electron–hole pairs. Duan et al. [99] successfully developed a CdS/sulfur-modified graphitic carbon nitride (GCNS) S-scheme heterojunction photocatalytic system based on solid-state diffusion. The experimental results demonstrated that, when the mass ratio of CdS to GCNS was 1:2, the composite catalyst achieved a 100.0% degradation efficiency for MO within 60 min of visible light irradiation. The apparent kinetic constant k was enhanced by factors of 9.67 and 5.39 compared to single-component GCNS and CdS, respectively. Through DFT simulations combined with carrier trajectory analysis, the study revealed the physical mechanism underlying directional electron migration at the heterointerface: unidirectional electron transfer from CdS to GCNS was attributed to the built-in electric field formed by band bending at the interface. This one-dimensional electron transport channel not only broadened the visible light absorption range but also significantly improved the generation efficiency of reactive oxygen species (especially •O2⁻) via the space charge separation effect. Liu et al. [100] innovatively constructed a g-C3N4/TiO2/ZnIn2S4 composite photocatalytic system (CTZA) featuring a double S-scheme heterojunction through an isoelectric-point-controlled calcination strategy, using a three-dimensional graphene aerogel as a structural template. As shown in Figure 6c, compared to traditional composites, CTZ components were uniformly distributed on the surface of graphene sheets, resulting in an ordered three-dimensional interconnected network with multi-level pore characteristics. This unique topological structure originated from the directional self-assembly process induced by π-π stacking interactions between reduced graphene oxide (rGO) layers [101]. This mechanism not only preserved the macroscopic integrity of the aerogel but also enabled the precise spatial arrangement of nanoscale components. High-resolution transmission electron microscopy characterization further revealed that a strong interface coupling effect was established between CTZ nano-units and the graphene aerogel substrate. The characteristic crystal plane spacings of 0.325 nm, 0.352 nm, and 0.191 nm observed in Figure 6d corresponded to the (002) lattice of g-C3N4, the (101) plane of anatase TiO2, and the (110) plane of hexagonal ZnIn2S4, respectively. These precise lattice parameter matches confirmed the successful construction of the ternary heterojunction at the atomic level. This ordered assembly ensured the structural stability of each component while constructing an efficient interfacial carrier transport pathway. Through electron spin resonance (ESR) spectroscopy analysis (Figure 6e-f), using DMPO as a radical scavenger, the study demonstrated that the CTZA composite material could simultaneously generate •OH and •O2⁻ under simulated solar light excitation. Experimental data showed that the characteristic signal peaks of DMPO-•OH and DMPO-•O2⁻ significantly increased under light conditions. Specifically, the •OH generation intensity of the CTZA system was markedly higher than that of the single TiO2 system, and the •O2⁻ generation efficiency exceeded those of single-component g-C3N4 and ZnIn2S4 systems. Coupling theoretical analysis with experimental data further confirmed that the construction of the double S-scheme heterojunction established a directional charge transport channel in the CTZA composite system. This synergistic mechanism enabled the material to achieve a 97.5% degradation efficiency for MO within 30 min of simulated solar light irradiation and a 98.3% reduction rate for Cr(VI) within 70 min, with the catalytic performance exceeding that of the single-component systems by more than two orders of magnitude. Notably, the three-dimensional porous structure of the material not only enhanced the removal of pollutants through physical adsorption but also enhanced the interfacial kinetics of the photocatalytic reaction by shortening carrier migration paths. This adsorption–catalysis synergy provides a new design concept for developing highly efficient environmental remediation materials. This study elucidated the carrier transport mechanism of the double S-scheme heterojunction within the CTZA composite photocatalytic system. As shown in Figure 6g, under photoexcitation conditions, photogenerated electrons in the CB of TiO2 recombine with holes in the VBs of g-C3N4 and ZnIn2S4 via interfacial recombination pathways. This dynamic equilibrium maintains the strong oxidation capability of VB holes in TiO2 while preserving the high reduction activity of CB electrons in g-C3N4 and ZnIn2S4. rGO, with its extended specific surface area and π-conjugated framework, plays a multifunctional role as an electron transport medium in the system. On the one hand, it acts as an efficient electron acceptor to promote exciton dissociation and directional charge migration [102,103]; on the other hand, it accelerates the interfacial transfer of CB electrons in g-C3N4 and ZnIn2S4 through its delocalized π-electron network, thereby significantly enhancing the spatial separation efficiency of photogenerated carriers. Both experiments and theoretical calculations confirm that this charge transport mechanism enables accumulated high-concentration electrons and holes to participate in molecular oxygen activation and water molecule oxidation, respectively, generating highly oxidative •OH and •O2⁻ radicals [104,105]. These reactive oxygen species effectively degrade organic pollutants such as MO into harmless small-molecule products (e.g., H2O and CO2). The core advantage lies in the synergistic enhancement of the double S-scheme heterojunction and the three-dimensional aerogel framework—where the former optimizes band structures to suppress carrier recombination and enhance redox capabilities, while the latter provides an ideal microenvironment and mass transfer channels for interfacial reactions.
Figure 6. (a) The influence of different capture agents on the photocatalytic degradation of MO by g-C3N4/ZnO/TiO2/Cork [96]. (b) The preparation process of the CoTiO3/g-C3N4 nanocomposite [97]. (c) An SEM image of the CTZA photocatalyst; (d) an HRTEM image of the CTZA photocatalyst; ESR spectra of (e) DMPO-•O2 and (f) DMPO-•OH for different photocatalysts; (g) a schematic diagram of the possible photocatalytic mechanism of the CTZA photocatalyst [100].
Figure 6. (a) The influence of different capture agents on the photocatalytic degradation of MO by g-C3N4/ZnO/TiO2/Cork [96]. (b) The preparation process of the CoTiO3/g-C3N4 nanocomposite [97]. (c) An SEM image of the CTZA photocatalyst; (d) an HRTEM image of the CTZA photocatalyst; ESR spectra of (e) DMPO-•O2 and (f) DMPO-•OH for different photocatalysts; (g) a schematic diagram of the possible photocatalytic mechanism of the CTZA photocatalyst [100].
Catalysts 15 00592 g006
Table 3. The photocatalytic dye degradation performance of g-C3N4-based S-scheme heterojunctions.
Table 3. The photocatalytic dye degradation performance of g-C3N4-based S-scheme heterojunctions.
PhotocatalystSynthesis MethodAmount of CatalystLight SourceDyeAmount of DyeTime
(min)
Efficiency (%)Ref.
g-C3N4/Co/ZnOUltrasound, sol–gel20 mg300 W
Xenon lamp,
λ > 420 nm
CV50 mL, 15 ppm120 74.5[27]
Ag/AgI/g-C3N4Hydrothermal/Simulated sunlightCV/60100.0[48]
Ca@TiO2@g-C3N4Thermal polymerization,
Ultrasound
50 mg500W
Xenon lamp
MG100 mL, 30 mg/L6099.9[94]
C-CeO2/g-C3N4Hydrothermal50 mg solarMG100 mL, 30 mg/L15091.9[95]
g-C3N4/ZnO/TiO2/CorkCo-precipitation60 mg250 W
halogen lamp
MO100 mL, 1 × 10−5 M6098.3[96]
CoTiO3/g-C3N4In situ calcination/Visible,
λ > 420 nm
MO/24099.7[97]
g-C3N4/Bi2S3/CuSHydrothermal10 mg VisibleMO100 mL, 0.1 mg/L6098.0[98]
CdS/GCNSSolid-state diffusion/VisibleMO/60100.0[99]
CTZAIsoelectric point calcination/Simulated sunlightMO/3097.5[100]
g-C3N4/Ag2WO4/Bi2S3In situ growth50 mg 140 W LED lampCR250 mL, 20 mg/L6098.0[106]
R.palustris/RCM@CPUHydrothermal/100 W lightCR50 mg/L48099.5[107]
V2O5/Ndef-g-C3N4Controllable pyrolysis5 mg18 W LED lampIC25 mL, 20 ppm3598.2[108]
NiMn2O4/g-C3N4Ultrasonic co-precipitation50 mg400 W lamp (Osram, Munich, Germany)EBT50 mL, 10 ppm12096.4[109]
MgO-TiO2@g-C3N4Ultrasound50 mg500 W
Xenon lamp,
λ > 420 nm
ARS100 mL, 30 mg/L6094.0[110]
CaSnO3/g-C3N4Ultrasonic co-precipitation45 mg400 W lamp (Osram, Munich, Germany)ER45 mL, 10 ppm 9086.2[111]

4.2.2. Congo Red

Jabbar et al. [106] successfully fabricated a novel S-scheme g-C3N4/Ag2WO4/Bi2S3 heterojunction via a multi-step synthesis strategy. The ternary photocatalyst achieved 98.0% degradation of Congo red under visible light irradiation within 60 min, as shown in Table 3. Theoretical analysis revealed that the ternary system realized the directional migration and spatial separation of photogenerated electron–hole pairs by constructing dual S-scheme charge transfer channels. Specifically, the conduction-band electrons of Ag2WO4 preferentially recombined with the valence-band holes of Bi2S3, while a secondary S-scheme transfer pathway was established between g-C3N4 and Bi2S3. This synergistic mechanism not only significantly suppressed carrier recombination but also preserved the intrinsic redox potential of each component through optimized band alignment. In the field of bio-photocatalytic synergy technology, Liu et al. [107] innovatively integrated an S-scheme rod-shaped g-C3N4/MoS2 heterojunction (RCM), prepared via hydrothermal synthesis, with a metabolically functional bacterial community. The RCM photocatalyst was immobilized onto the surface of a chitosan-modified polyurethane sponge (CPU) through surface modification techniques, while Rhodopseudomonas palustris, a broad-spectrum metabolic bacterium, was encapsulated within the carrier matrix. The resulting composite system (R. palustris/RCM@CPU) achieved a removal efficiency of 99.5% for Congo red under simulated solar light irradiation over 8 h. Its superior performance can be attributed to a triple synergistic effect: firstly, the S-scheme band structure of RCM enabled efficient charge separation through the interfacial built-in electric field; secondly, R. palustris exhibited excellent environmental adaptability and could mineralize photocatalytic intermediates via enzymatic catalysis pathways; finally, the chitosan-modified CPU carrier demonstrated significantly enhanced surface roughness and functional group density, which not only improved the mechanical stability of the photocatalytic layer but also provided an ideal microecological niche for microbial colonization. This cross-scale “photocatalysis–biodegradation” coupling mechanism offers an innovative solution for the treatment of refractory organic pollutants.

4.2.3. Other Anionic Dyes

Hassan et al. [108] successfully developed a vanadium pentoxide/nitrogen-deficient graphitic carbon nitride (VO/Ndef-CN) composite photocatalytic system via a controllable pyrolysis strategy. The 5VO/Ndef-CN heterostructure achieved a degradation efficiency of 98.2% for indigo carmine dye within 35 min under visible light irradiation, as shown in Table 3. The performance enhancement mechanism primarily originates from two aspects: (1) the S-scheme heterojunction establishes a directional carrier transport channel through band bending, effectively suppressing the recombination of photogenerated electron–hole pairs; (2) nitrogen defect engineering in conjunction with the heterointerface significantly broadens the material’s light response range and enhances its photon capture capability. Yaqoubi et al. [109] synthesized a NiMn2O4/g-C3N4 heterojunction using an ultrasonic co-precipitation method. In an acidic medium, the NiMn2O4/g-C3N4 nanocomposite degraded 96.4% of chrome black T within 120 min under visible light irradiation. The S-scheme heterojunction facilitated the efficient separation of photogenerated electron–hole pairs. Alqarni et al. [110] fabricated a ternary MgO-TiO2@g-C3N4 composite through a facile ultrasonic synthesis method. The MgO-TiO2@g-C3N4 heterojunction degraded 94.0% of alizarin red S within 60 min under visible light irradiation. The S-scheme heterojunction spatially separated and retained high concentrations of h⁺ and e⁻ at the VB and CB of g-C3N4 and MgO-TiO2, thereby enabling effective dye degradation. Hosseini et al. [111] prepared a CaSnO3/g-C3N4 nanocomposite via an ultrasonic-assisted co-precipitation method. The CaSnO3/g-C3N4 nanocomposite decomposed 86.2% of erythrosine within 90 min under visible light irradiation. Owing to the efficient utilization of visible light and enhanced charge carrier separation by the S-scheme heterojunction, this nanocomposite demonstrates excellent performance in degrading organic dyes.
In summary, innovative design strategies have been used to develop S-scheme heterojunction g-C3N4 structures with remarkable photocatalytic performance in the degradation of anionic dyes. Through various advanced preparation techniques, including co-precipitation, physical blending, in situ calcination, solvothermal synthesis, and solid-state diffusion, researchers have successfully developed a series of g-C3N4-based S-scheme heterojunction photocatalytic systems. These systems effectively suppress the recombination of photogenerated electron–hole pairs and enhance the spectral response range through optimized band structure engineering and carrier migration mechanisms while significantly improving the generation efficiency of reactive oxygen species. In degradation experiments involving methyl orange, Congo red, and other anionic dyes, these S-scheme heterojunction composites consistently demonstrated high removal efficiency and superior photocatalytic performance. These research findings not only provide novel insights into the rational design of efficient photocatalytic materials but also offer innovative solutions for addressing refractory organic pollutants in environmental remediation.

5. Conclusions

With the continuous discharge of dye-containing toxic wastewater in modern agricultural and industrial production processes, environmental remediation technologies based on photocatalysis have demonstrated significant application value in industrial wastewater treatment due to their high mineralization efficiency, environmental friendliness, and energy economy. Compared with traditional wide-bandgap semiconductor materials such as TiO2, which suffer from inherent limitations, including a weak visible light response and low quantum efficiency, S-scheme heterojunction photocatalysts developed in recent years offer an innovative solution to overcome the charge recombination bottleneck of conventional type-II and Z-scheme heterojunctions. By constructing a stepped band structure, these catalysts enable the efficient separation of photogenerated carriers while preserving the redox potential of each component. Graphitic carbon nitride, characterized by its narrow band gap, excellent visible light absorption, and superior physicochemical stability, has emerged as an ideal candidate for constructing S-scheme heterojunctions. Its two-dimensional layered structure facilitates charge transfer and significantly enhances photocatalytic activity. Studies have shown that g-C3N4-based S-scheme heterojunctions exhibit remarkable photocatalytic performance in degrading various cationic dyes (e.g., MB, RhB, and crystal violet) and anionic dyes (e.g., Congo red and MO), achieving degradation efficiencies as high as nearly 100.0% in some cases. Researchers have successfully fabricated diverse g-C3N4-based S-scheme heterojunction systems using various synthesis methods, including MoS2/g-C3N4, CdS/g-C3N4, and PbTiO3/g-C3N4, among others, all of which demonstrate excellent photocatalytic degradation capabilities. Furthermore, this review systematically discusses the performance differences and underlying mechanisms of different catalyst systems in degrading various types of dyes.

6. Prospects

Significant advancements have been achieved in the application of g-C3N4-based S-scheme heterojunctions for the photocatalytic degradation of dye wastewater. However, their practical implementation still faces numerous challenges. Future research can focus on the following directions to promote the industrialization and performance optimization of this technology.

6.1. Development and Optimization of Novel Heterojunction Systems

In the field of semiconductor coupling system innovation, more semiconductor materials compatible with g-C3N4 (BiOX, MXene, COF, etc.) should be explored to construct advanced S-scheme heterojunctions, addressing the limitations of light absorption and charge separation efficiency in existing systems. For example, introducing materials with broad spectral responses (e.g., black phosphorus) or magnetic components (e.g., Fe3O4) can enhance light absorption and enable convenient catalyst recovery. To improve interfacial charge transfer efficiency, surface modification strategies such as doping, defect engineering, and cocatalyst loading can be employed. Specifically, doping with metal single atoms (e.g., Pt or Co) or non-metal elements (e.g., P or B) can increase the density of active sites and extend the lifetime of photogenerated carriers. At the microstructure engineering level, three-dimensional porous g-C3N4-based heterojunctions (e.g., core–shell structures, nanoflower-like structures) can be developed to maximize the specific surface area and exposure of reactive sites, thereby enhancing the adsorption and degradation kinetics of dye molecules.

6.2. Coupling of Multiple Technologies and Intelligent Control

In the domain of energy–mass synergy catalysis, a photocatalytic–electrocatalytic coupling system can be constructed to accelerate charge separation via an external electric field while simultaneously generating oxidants such as H2O2 through water electrolysis to enhance degradation efficiency. Additionally, in the direction of photothermal regulation, the synergistic mechanism between photocatalysis and thermal catalysis can be investigated. For instance, incorporating photothermal materials (e.g., Au or Cu2O) can leverage the photothermal effect to elevate reaction temperatures and expedite dye molecule decomposition rates. Regarding intelligent technology integration, machine learning models can be utilized to predict optimal heterojunction compositions and synthesis conditions, reducing experimental cycles. Real-time monitoring of wastewater treatment effects using sensors can enable intelligent control of the catalytic system. In terms of process coupling innovation, the feasibility of integrating photocatalysis with membrane separation, biological degradation, and other technologies can be explored to establish multi-stage treatment processes, improving the overall wastewater treatment efficiency and economic viability.
Future research should prioritize the core objectives of “efficiency” and “intelligence”, driving the transition of g-C3N4-based S-scheme heterojunctions from laboratory-scale studies to industrial applications through material innovation and interdisciplinary collaboration. This process not only involves addressing scientific challenges but also requires consideration of environmental safety and economic feasibility, ultimately providing sustainable technical solutions for global dye wastewater management.

Author Contributions

X.S.: investigation, writing—original draft preparation, review and editing; Z.M.: methodology, visualization; Z.W.: methodology; S.J.: visualization; J.H.: editing; P.X.: supervision; Y.C.: writing—original draft preparation; All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Raw data are available upon request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Elumalai, P.; Gao, X.; Parthipan, P.; Luo, J.; Cui, J. Agrochemical pollution: A serious threat to environmental health. Curr. Opin. Environ. Sci. Health 2025, 43, 100597. [Google Scholar] [CrossRef]
  2. Ismail, M.; Akhtar, K.; Khan, M.; Kamal, T.; Khan, M.A.; Asiri, A.M.; Seo, J.; Khan, S.B. Pollution, toxicity and carcinogenicity of organic dyes and their catalytic bio-remediation. Curr. Pharm. Des. 2019, 25, 3645–3663. [Google Scholar] [CrossRef]
  3. Lima, J.P.; Alvarenga, G.; Goszczynski, A.C.; Rosa, G.R.; Lopes, T.J. Batch adsorption of methylene blue dye using Enterolobium contortisiliquum as bioadsorbent: Experimental, mathematical modeling and simulation. J. Ind. Eng. Chem. 2020, 91, 362–371. [Google Scholar] [CrossRef]
  4. Naddeo, V.; Secondes, M.F.N.; Borea, L.; Hasan, S.W.; Ballesteros, F., Jr.; Belgiorno, V. Removal of contaminants of emerging concern from real wastewater by an innovative hybrid membrane process—UltraSound, Adsorption, and Membrane ultrafiltration. Ultrason. Sonochem. 2020, 68, 105237. [Google Scholar] [CrossRef] [PubMed]
  5. Geed, S.R.; Samal, K.; Tagade, A.J. Development of adsorption-biodegradation hybrid process for removal of methylene blue from wastewater. J. Environ. Chem. Eng. 2019, 7, 103439. [Google Scholar] [CrossRef]
  6. Rodriguez, N.O.M.; Picos, A.R.; Bravo, Y.N.; Pacheco, A.M.; Martínez, H.C.A.; Peralta-Hernández, J.M. Electrochemical oxidation technology to treat textile wastewaters. Curr. Opin. Electrochem. 2021, 29, 100806. [Google Scholar] [CrossRef]
  7. Liu, C.; Dong, S.; Chen, Y. Enhancement of visible-light-driven photocatalytic activity of carbon plane/g-C3N4/TiO2 nanocomposite by improving heterojunction contact. Chem. Eng. J. 2019, 371, 706–718. [Google Scholar] [CrossRef]
  8. Do, H.T.; Phan Thi, L.A.; Dao Nguyen, N.H.; Huang, C.W.; Le, Q.V.; Nguyen, V.H. Tailoring photocatalysts and elucidating mechanisms of photocatalytic degradation of perfluorocarboxylic acids (PFCAs) in water: A comparative overview. J. Chem. Technol. Biotechnol. 2020, 95, 2569–2578. [Google Scholar] [CrossRef]
  9. Xie, Z.; Saad, A.; Shang, Y.; Wang, Y.; Luo, S.; Wei, Z. Enhanced degradation of micropollutants by visible light photocatalysts with strong oxygen activation ability. Water Res. 2023, 247, 120785. [Google Scholar] [CrossRef]
  10. Sacco, O.; Mancuso, A.; Venditto, V.; Pragliola, S.; Vaiano, V. Behavior of N-doped TiO2 and N-doped ZnO in photocatalytic azo dye degradation under UV and visible light irradiation: A preliminary investigation. Catalysts 2022, 12, 1208. [Google Scholar] [CrossRef]
  11. Sun, F.; Qi, H.; Xie, Y.; Ma, Q.; He, W.; Xu, D.; Wang, G.; Yu, W.; Wang, T.; Dong, X. Flexible self-supporting bifunctional [TiO2/C]//[Bi2WO6/C] carbon-based Janus nanofiber heterojunction photocatalysts for efficient hydrogen evolution and degradation of organic pollutant. J. Alloys Compd. 2020, 830, 154673. [Google Scholar] [CrossRef]
  12. Zia, J.; Riaz, U. Photocatalytic degradation of water pollutants using conducting polymer-based nanohybrids: A review on recent trends and future prospects. J. Mol. Liq. 2021, 340, 117162. [Google Scholar] [CrossRef]
  13. Li, X.; Kang, B.; Dong, F.; Zhang, Z.; Luo, X.; Han, L.; Huang, J.; Feng, Z.; Chen, Z.; Xu, J.; et al. Enhanced photocatalytic degradation and H2/H2O2 production performance of S-pCN/WO2.72 S-scheme heterojunction with appropriate surface oxygen vacancies. Nano Energy 2021, 81, 105671. [Google Scholar] [CrossRef]
  14. Du, H.; Liu, Y.N.; Shen, C.C.; Xu, A.W. Nanoheterostructured photocatalysts for improving photocatalytic hydrogen production. Chin. J. Catal. 2017, 38, 1295–1306. [Google Scholar] [CrossRef]
  15. Jiang, G.; Zheng, C.; Yan, T.; Jin, Z. Cd0.8Mn0.2S/MoO3 composites with an S-scheme heterojunction for efficient photocatalytic hydrogen evolution. Dalton Trans. 2021, 50, 5360–5369. [Google Scholar] [CrossRef]
  16. Li, Y.; Zhou, M.; Cheng, B.; Shao, Y. Recent advances in g-C3N4-based heterojunction photocatalysts. J. Mater. Sci. Technol. 2020, 56, 1–17. [Google Scholar] [CrossRef]
  17. Zhang, L.; Zhang, J.; Yu, H.; Yu, J. Emerging S-scheme photocatalyst. Adv. Mater. 2022, 34, 2107668. [Google Scholar] [CrossRef]
  18. Preeyanghaa, M.; Vinesh, V.; Neppolian, B. Construction of S-scheme 1D/2D rod-like g-C3N4/V2O5 heterostructure with enhanced sonophotocatalytic degradation for Tetracycline antibiotics. Chemosphere 2022, 287, 132380. [Google Scholar] [CrossRef]
  19. Wang, L.; Cheng, B.; Zhang, L.; Yu, J. In situ irradiated XPS investigation on S-scheme TiO2@ZnIn2S4 photocatalyst for efficient photocatalytic CO2 reduction. Small 2021, 17, 2103447. [Google Scholar] [CrossRef]
  20. Feng, K.; Tian, J.; Hu, X.; Fan, J.; Liu, E. Active-center-enriched Ni0.85Se/g-C3N4 S-scheme heterojunction for efficient photocatalytic H2 generation. Int. J. Hydrogen Energy 2022, 47, 4601–4613. [Google Scholar] [CrossRef]
  21. Chen, Y.; Wang, Q.; Huang, H.; Kou, J.; Lu, C.; Xu, Z. Effective solar driven H2 production by Mn0.5Cd0.5Se/g-C3N4 S-scheme heterojunction photocatalysts. Int. J. Hydrogen Energy 2021, 46, 32514–32522. [Google Scholar] [CrossRef]
  22. Wang, L.; Chuanbiao, B.; Yu, J. Challenges of Z-scheme photocatalytic mechanisms. Trends Chem. 2022, 4, 973–983. [Google Scholar] [CrossRef]
  23. Han, T.; Shi, H.; Chen, Y. Facet-dependent CuO/{010} BiVO4 S-scheme photocatalyst enhanced peroxymonosulfate activation for efficient norfloxacin removal. J. Mater. Sci. Technol. 2024, 174, 30–43. [Google Scholar] [CrossRef]
  24. Chen, R.; Xia, J.; Chen, Y.; Shi, H. S-scheme-enhanced PMS activation for rapidly degrading tetracycline using CuWO4−x/Bi12O17Cl2 heterostructures. Acta Phys. Chim. Sin. 2023, 39, 2209012. [Google Scholar] [CrossRef]
  25. Li, Z.; Wu, Z.; He, R.; Wan, L.; Zhang, S. In2O3-x(OH)y/Bi2MoO6 S-scheme heterojunction for enhanced photocatalytic performance. J. Mater. Sci. Technol. 2020, 56, 151–161. [Google Scholar] [CrossRef]
  26. Zhang, B.; Hu, X.; Liu, E.; Fan, J. Novel S-scheme 2D/2D BiOBr/g-C3N4 heterojunctions with enhanced photocatalytic activity. Chin. J. Catal. 2021, 42, 1519–1529. [Google Scholar] [CrossRef]
  27. Leelavathi, H.; Muralidharan, R.; Abirami, N.; Tamizharasan, S.; Sankeetha, S.; Kumarasamy, A.; Arulmozhi, R. Construction of step-scheme g-C3N4/Co/ZnO heterojunction photocatalyst for aerobic photocatalytic degradation of synthetic wastewater. Colloids Surf. A 2023, 656, 130449. [Google Scholar] [CrossRef]
  28. Le, S.; Zhu, C.; Cao, Y.; Wang, P.; Liu, Q.; Zhou, H.; Chen, C.; Wang, S.; Duan, X. V2O5 nanodot-decorated laminar C3N4 for sustainable photodegradation of amoxicillin under solar light. Appl. Catal. B Environ. 2022, 303, 120903. [Google Scholar] [CrossRef]
  29. Ren, X.; Guo, M.; Xue, L.; Xu, L.; Li, L.; Yang, L.; Wang, M.; Xin, Y.; Ding, F.; Wang, Y. Photoelectrochemical performance and S-scheme mechanism of ternary GO/g-C3N4/TiO2 heterojunction photocatalyst for photocatalytic antibiosis and dye degradation under visible light. Appl. Surf. Sci. 2023, 630, 157446. [Google Scholar] [CrossRef]
  30. Chen, X.; Wang, J.; Chai, Y.; Zhang, Z.; Zhu, Y. Efficient photocatalytic overall water splitting induced by the giant internal electric field of a g-C3N4/rGO/PDIP Z-scheme heterojunction. Adv. Mater. 2021, 33, 2007479. [Google Scholar] [CrossRef]
  31. Cheng, C.; He, B.; Fan, J.; Cheng, B.; Cao, S.; Yu, J. An inorganic/organic S-scheme heterojunction H2-production photocatalyst and its charge transfer mechanism. Adv. Mater. 2021, 33, 2100317. [Google Scholar] [CrossRef]
  32. Cheng, L.; Zhang, H.; Li, X.; Fan, J.; Xiang, Q. Carbon-graphitic carbon nitride hybrids for heterogeneous photocatalysis. Small 2021, 17, 2005231. [Google Scholar] [CrossRef]
  33. Jiang, Z.; Zhang, X.; Chen, H.S.; Yang, P.; Jiang, S.P. Fusiform-Shaped g-C3N4 Capsules with Superior Photocatalytic Activity. Small 2020, 16, 2003910. [Google Scholar] [CrossRef]
  34. Yi, J.; El-Alami, W.; Song, Y.; Li, H.; Ajayan, P.M.; Xu, H. Emerging surface strategies on graphitic carbon nitride for solar driven water splitting. Chem. Eng. J. 2020, 382, 122812. [Google Scholar] [CrossRef]
  35. Hsini, A.; Naciri, Y.; Benafqir, M.; Ajmal, Z.; Aarab, N.; Laabd, M.; Navío, J.; Puga, F.; Boukherroub, R.; Bakiz, B.; et al. Facile synthesis and characterization of a novel 1,2,4,5-benzene tetracarboxylic acid doped polyaniline@zinc phosphate nanocomposite for highly efficient removal of hazardous hexavalent chromium ions from water. J. Colloid Interface Sci. 2021, 585, 560–573. [Google Scholar] [CrossRef]
  36. Yang, W.; Kim, J.H.; Hutter, O.S.; Phillips, L.J.; Tan, J.; Park, J.; Lee, H.; Major, J.D.; Lee, J.S.; Moon, J.; et al. Benchmark performance of low-cost Sb2Se3 photocathodes for unassisted solar overall water splitting. Nat. Commun. 2020, 11, 861. [Google Scholar] [CrossRef]
  37. Kumar, P.; Laishram, D.; Sharma, R.K.; Vinu, A.; Hu, J.; Kibria, M.G. Boosting photocatalytic activity using carbon nitride based 2D/2D van der Waals heterojunctions. Chem. Mater. 2021, 33, 9012–9092. [Google Scholar] [CrossRef]
  38. Yaqoubi, M.; Salavati-Niasari, M.; Ghanbari, M. S-Scheme CuMn2O4/g-C3N4 heterojunction: Fabrication, characterization, and investigation of photodegradation potential of organic pollutants. Appl. Water Sci. 2025, 15, 13. [Google Scholar] [CrossRef]
  39. Tran Huu, H.; Thi, M.D.N.; Nguyen, V.P.; Thi, L.N.; Phan, T.T.T.; Hoang, Q.D.; Luc, H.H.; Kim, S.J.; Vo, V. One-pot synthesis of S-scheme MoS2/g-C3N4 heterojunction as effective visible light photocatalyst. Sci. Rep. 2021, 11, 14787. [Google Scholar] [CrossRef]
  40. Li, Y.; Xia, Z.; Yang, Q.; Wang, L.; Xing, Y. Review on g-C3N4-based S-scheme heterojunction photocatalysts. J. Mater. Sci. Technol. 2022, 125, 128–144. [Google Scholar] [CrossRef]
  41. Li, Q.; Zhao, W.; Zhai, Z.; Ren, K.; Wang, T.; Guan, H.; Shi, H. 2D/2D Bi2MoO6/g-C3N4 S-scheme heterojunction photocatalyst with enhanced visible-light activity by Au loading. J. Mater. Sci. Technol. 2020, 56, 216–226. [Google Scholar] [CrossRef]
  42. Liu, Q.; He, X.; Peng, J.; Yu, X.; Tang, H.; Zhang, J. Hot-electron-assisted S-scheme heterojunction of tungsten oxide/graphitic carbon nitride for broad-spectrum photocatalytic H2 generation. Chin. J. Catal. 2021, 42, 1478–1487. [Google Scholar] [CrossRef]
  43. Guo, X.; Xi, Y.; Li, Y.; Zhu, J.; Yan, H.; Zha, F.; Tang, X.; Tian, H. Ag regulates the interfacial electric field of g-C3N4/FeVO4 S-scheme heterojunction for activated H2O2 degradation of organic pollutants in complex environments. Chem. Eng. J. 2025, 505, 159640. [Google Scholar] [CrossRef]
  44. Dai, Z.; Zhen, Y.; Sun, Y.; Li, L.; Ding, D. ZnFe2O4/g-C3N4 S-scheme photocatalyst with enhanced adsorption and photocatalytic activity for uranium(VI) removal. Chem. Eng. J. 2021, 415, 129002. [Google Scholar] [CrossRef]
  45. Chu, Z.; Li, J.; Sohn, H.Y.; Chen, C.; Huang, X.; Lan, Y.; Murali, A.; Zhang, J. CeO2-g-C3N4 S-scheme heterojunctions for enhanced photocatalytic performance: Effects of surface C/N ratio on photocatalytic and adsorption properties. Compos. Part B Eng. 2023, 257, 110689. [Google Scholar] [CrossRef]
  46. Yang, C.; Yang, J.; Liu, S.; Zhao, M.; Duan, X.; Wu, H.; Liu, L.; Liu, W.; Li, J.; Ren, S. Constructing C–O bridged CeO2/g-C3N4 S-scheme heterojunction for methyl orange photodegradation: Experimental and theoretical calculation. J. Environ. Manag. 2023, 335, 117608. [Google Scholar] [CrossRef]
  47. Chen, P.; Zhu, Z.; Liu, Z.; Liang, F.; Zhu, X.; Bin, Z.; Huang, F.; Wang, N.; Zhu, Y. Efficient removal of ciprofloxacin from water by BiOX/GaMOF S-scheme heterojunction: A synergistic effect of adsorption and photocatalysis. Chem. Eng. J. 2025, 506, 159689. [Google Scholar] [CrossRef]
  48. Sun, F.; Xu, Q.; Liu, H.; Xu, D.; Wang, X.; Luo, C.; Wang, T.; Yu, H.; Yu, W.; Dong, X. Plasmon and N-vacancy synergistically enhanced tubular carbon nitride-based S-scheme heterojunction photocatalyst with one stone five birds function: Pathways, DFT calculation and mechanism insight. J. Catal. 2024, 440, 115813. [Google Scholar] [CrossRef]
  49. Cheng, C.; Zhang, J.; Zhu, B.; Liang, G.; Zhang, L.; Yu, J. Verifying the charge-transfer mechanism in S-Scheme heterojunctions using femtosecond transient absorption spectroscopy. Angew. Chem. Int. Ed. 2023, 62, e202218688. [Google Scholar] [CrossRef]
  50. Guo, N.; Zeng, Y.; Li, H.; Xu, X.; Yu, H.; Han, X. Novel mesoporous TiO2@g-C3N4 hollow core@shell heterojunction with enhanced photocatalytic activity for water treatment and H2 production under simulated sunlight. J. Hazard. Mater. 2018, 353, 80–88. [Google Scholar] [CrossRef]
  51. Yan, Q.; Fu, Y.; Zhang, Y.; Wang, H.; Wang, S.; Cui, W. Ag/γ-AgI/Bi2O2CO3/Bi S-scheme heterojunction with enhanced photocatalyst performance. Sep. Purif. Technol. 2021, 263, 118389. [Google Scholar] [CrossRef]
  52. Lee, Y.; Jeong, Y.J.; Cho, I.S.; Park, S.; Lee, C.; Alvarez, P.J. Facile synthesis of N vacancy g-C3N4 using Mg-induced defect on the amine groups for enhanced photocatalytic •OH generation. J. Hazard. Mater. 2023, 449, 131046. [Google Scholar] [CrossRef] [PubMed]
  53. Sun, J.; Zhang, B.; Chen, W.; Tao, Z.; Liu, J.; Wang, L. Biochar doped carbon nitride to enhance the photocatalytic hydrogen evolution through synergy of nitrogen vacancies and bridging carbon structure: Nanoarchitectonics and first-principles calculation. Carbon 2023, 209, 117988. [Google Scholar] [CrossRef]
  54. Fang, H.; Guo, H.; Niu, C.; Liang, C.; Huang, D.-W.; Tang, N.; Liu, H.; Yang, Y.; Li, L. Hollow tubular graphitic carbon nitride catalyst with adjustable nitrogen vacancy: Enhanced optical absorption and carrier separation for improving photocatalytic activity. Chem. Eng. J. 2020, 402, 126185. [Google Scholar] [CrossRef]
  55. Venkatesh, G.; Palanisamy, G.; Srinivasan, M.; Vignesh, S.; Elavarasan, N.; Pazhanivel, T.; Al-Enizi, A.M.; Ubaidullah, M.; Karim, A.; Prabu, K.M. CaSnO3 coupled g-C3N4 S-scheme heterostructure photocatalyst for efficient pollutant degradation. Diam. Relat. Mater. 2022, 124, 108873. [Google Scholar] [CrossRef]
  56. Patra, R.; Panda, P.K.; Lin, T.; Wu, M.; Yang, P. Graphitic carbon nitride nanosheet and ferroelectric PbTiO3 nanoplates S-scheme heterostructure for enhancing hydrogen production and textile dye degradation. Chem. Eng. Sci. 2024, 295, 120133. [Google Scholar] [CrossRef]
  57. Abedini, E.; Roudgar-Amoli, M.; Alizadeh, A.; Shariatinia, Z. S-scheme heterojunctions based on novel Sm2CeMnO6 double perovskite oxide and g-C3N4 with excellent photocatalytic dye degradation performances. Environ. Sci. Pollut. Res. 2023, 30, 114956–114984. [Google Scholar] [CrossRef]
  58. Wang, J.; Ren, P.; Du, Y.; Zhao, X.; Chen, Z.; Pei, L.; Jin, Y. Construction of tubular g-C3N4/TiO2 S-scheme photocatalyst for high-efficiency degradation of organic pollutants under visible light. J. Alloys Compd. 2023, 947, 169659. [Google Scholar] [CrossRef]
  59. Wang, W.; Zeng, Z.; Zeng, G.; Zhang, C.; Xiao, R.; Zhou, C.; Xiong, W.; Yang, Y.; Lei, L.; Liu, Y.; et al. Sulfur doped carbon quantum dots loaded hollow tubular g-C3N4 as novel photocatalyst for destruction of Escherichia coli and tetracycline degradation under visible light. Chem. Eng. J. 2019, 378, 122132. [Google Scholar] [CrossRef]
  60. Barzegar, M.H.; Sabzehmeidani, M.M.; Ghaedi, M.; Avargani, V.M.; Moradi, Z.; Roy, V.A.L.; Heidari, H. S-scheme heterojunction g-C3N4/TiO2 with enhanced photocatalytic activity for degradation of a binary mixture of cationic dyes using solar parabolic trough reactor. Chem. Eng. Res. Des. 2021, 174, 307–318. [Google Scholar] [CrossRef]
  61. Riazati, P.; Sheibani, S. Enhancing photocatalytic water remediation by g-C3N4 through controlled CuO content in an S-scheme g-C3N4/CuO nanocomposite. J. Alloys Compd. 2025, 1016, 179008. [Google Scholar] [CrossRef]
  62. Ajami, A.; Sheibani, S.; Ataie, A. S-scheme CoFe2O4/g-C3N4 nanocomposite with high photocatalytic activity and antibacterial capability under visible light irradiation. J. Mater. Res. Technol. 2024, 30, 2168–2185. [Google Scholar] [CrossRef]
  63. Alsalme, A.; Alsaeedi, H.; Fayez, M.; Elmoneim, K.M.A.; Soltan, A.; Fahmy, M.; Ahmed, M.A. Sonochemical preparation of powerful S-scheme Zr(HPO4)2/g-C3N4 heterojunction for photocatalytic degradation of rhodamine B under natural solar radiations. J. Phys. Chem. Solids 2025, 196, 112392. [Google Scholar] [CrossRef]
  64. Chopan, N.A.; Chishti, H.T.N. Polypyrrole-decorated ZnO/g-C3N4 S-scheme photocatalyst for rhodamine B dye degradation: Mechanism and antibacterial activity. Mater. Today Chem. 2023, 32, 101643. [Google Scholar] [CrossRef]
  65. Shoaib, M.; Naz, M.Y.; Shukrullah, S.; Munir, M.A.; Irfan, M.; Rahman, S.; Ghanim, A.A. Dual S-Scheme Heterojunction CdS/TiO2/g-C3N4 Photocatalyst for Hydrogen Production and Dye Degradation Applications. ACS Omega 2023, 8, 43139–43150. [Google Scholar] [CrossRef]
  66. Alsalme, A.; Hassan, M.M.; Eltawil, M.A.; Amin, A.E.; Soltan, A.; Messih, M.F.A.; Ahmed, M.A. Rational sonochemical engineering of Ag2CrO4/g-C3N4 heterojunction for eradicating RhB dye under full broad spectrum. Heliyon 2024, 10, e31221. [Google Scholar] [CrossRef]
  67. Alsulmi, A.; Hussein, I.A.; Nasherty, M.; Hesham, M.; Soltan, A.; Messih, M.F.A.; Ahmed, M.A. Sonochemical Fabrication of S-Scheme AgI/g-C3N4 Heterojunction for Efficient Photocatalytic Degradation of RhB Dye. J. Inorg. Organomet. Polym. Mater. 2023, 34, 640–654. [Google Scholar] [CrossRef]
  68. Rajendran, R.; Vignesh, S.; Suganthi, S.; Raj, V.; Kavitha, G.; Palanivel, B.; Shkir, M.; Algarni, H. g-C3N4/TiO2/CuO S-scheme heterostructure photocatalysts for enhancing organic pollutant degradation. J. Phys. Chem. Solids 2022, 161, 110391. [Google Scholar] [CrossRef]
  69. Alsulmi, A.; Mohammed, N.N.; Hassan, M.M.; Eltawil, M.A.; Amin, A.E.; Fahmy, M.; Sultan, A.; Ahmed, M.A. Rational engineering of S-scheme CeO2/g-C3N4 heterojunctions for effective photocatalytic destruction of rhodamine B dye under natural solar radiations. Colloids Surf. A 2024, 689, 133683. [Google Scholar] [CrossRef]
  70. Alsalme, A.; Hesham, M.; Soltan, A.; Mohammed, N.N.; Nejm, A.Q.; Messih, M.F.A.; Hussein, I.A.; Ahmed, M.A. Pioneering the design of S-scheme SnS2/g-C3N4 nanocomposites via sonochemical and physical mixing methods for solar degradation of cationic rhodamine B dye. Mater. Adv. 2024, 5, 7278–7295. [Google Scholar] [CrossRef]
  71. Alsulmi, A.; Shaker, M.H.; Basely, A.M.; Abdel-Messih, M.F.; Sultan, A.; Ahmed, M.A. Engineering S-scheme Ag2CO3/g-C3N4 heterojunctions sonochemically to eradicate Rhodamine B dye under solar irradiation. RSC Adv. 2023, 13, 12229–12243. [Google Scholar] [CrossRef] [PubMed]
  72. Huang, Q.; Zhu, F.; Xiao, F.; Zhang, G.; Hou, H.; Bi, J.; Yan, S.; Hao, H. Construction of S-scheme heterogeneous HKUST-1/g-C3N4 with the piezoelectric effect for enhanced piezo-photocatalytic performance. Solid State Sci. 2023, 144, 107303. [Google Scholar] [CrossRef]
  73. Basely, A.M.; Shaker, M.H.; Helmy, F.M.; Abdel-Messih, M.F.; Ahmed, M.A. Construction of Bi2S3/g-C3N4 step S-scheme heterojunctions for photothermal decomposition of rhodamine B dye under natural sunlight radiations. Inorg. Chem. Commun. 2023, 148, 110300. [Google Scholar] [CrossRef]
  74. Elavarasan, N.; Vignesh, S.; Srinivasan, M.; Venkatesh, G.; Palanisamy, G.; Ramasamy, P.; Palanivel, B.; Al-Enizi, A.M.; Ubaidullah, M.; Minnam Reddy, V.R.; et al. Synergistic S-Scheme mechanism insights of g-C3N4 and rGO combined ZnO-Ag heterostructure nanocomposite for efficient photocatalytic and anticancer activities. J. Alloys Compd. 2022, 906, 164255. [Google Scholar] [CrossRef]
  75. Rajendran, R.; Mani, A. Photocatalytic, antibacterial and anticancer activity of silver-doped zinc oxide nanoparticles. J. Saudi Chem. Soc. 2020, 24, 1010–1024. [Google Scholar] [CrossRef]
  76. Lee, S.J.; Begildayeva, T.; Jung, H.J.; Koutavarapu, R.; Yu, Y.; Choi, M.; Choi, M.Y. Plasmonic ZnO/Au/g-C3N4 nanocomposites as solar light active photocatalysts for degradation of organic contaminants in wastewater. Chemosphere 2021, 263, 128262. [Google Scholar] [CrossRef]
  77. Adhikari, S.P.; Pant, H.R.; Kim, J.H.; Kim, H.J.; Park, C.H.; Kim, C.S. One pot synthesis and characterization of Ag-ZnO/g-C3N4 photocatalyst with improved photoactivity and antibacterial properties. Colloids Surf. A 2015, 482, 477–484. [Google Scholar] [CrossRef]
  78. Sharma, K.; Sudhaik, A.; Raizada, P.; Thakur, P.; Pham, X.M.; Van Le, Q.; Nguyen, V.; Ahamad, T.; Thakur, S.; Singh, P. Constructing α-Fe2O3/g-C3N4/SiO2 S-scheme-based heterostructure for photo-Fenton like degradation of rhodamine B dye in aqueous solution. Environ. Sci. Pollut. Res. 2023, 30, 124902–124920. [Google Scholar] [CrossRef]
  79. Bui, D.; Nguyen, T.; Le Vo, T.T.; Cao, T.M.; You, S.; Pham, V.V. SnO2–x Nanoparticles Decorated on Graphitic Carbon Nitride as S-Scheme Photocatalysts for Activation of Peroxymonosulfate. ACS Appl. Nano Mater. 2021, 4, 9333–9343. [Google Scholar] [CrossRef]
  80. Mishra, S.; Acharya, L.; Sharmila, S.; Sanjay, K.; Acharya, R. Designing g-C3N4/NiFe2O4 S-scheme heterojunctions for efficient photocatalytic degradation of Rhodamine B and tetracycline hydrochloride. Appl. Surf. Sci. Adv. 2024, 24, 100647. [Google Scholar] [CrossRef]
  81. Liang, H.; Zhao, J.; Brouzgou, A.; Wang, A.; Jing, S.; Kannan, P.; Chen, F.; Tsiakaras, P. Efficient photocatalytic H2O2 production and photodegradation of RhB over K-doped g-C3N4/ZnO S-scheme heterojunction. J. Colloid Interface Sci. 2025, 677, 1120–1133. [Google Scholar] [CrossRef] [PubMed]
  82. Bi, K.; Qin, X.; Cheng, S.; Liu, S. ZnO/g-C3N4 S scheme photocatalytic material with visible light response and enhanced photocatalytic performance. Diam. Relat. Mater. 2023, 137, 110143. [Google Scholar] [CrossRef]
  83. Majdoub, M.; Anfar, Z.; Amedlous, A. Emerging chemical functionalization of g-C3N4: Covalent/noncovalent modifications and applications. ACS Nano 2020, 14, 12390–12469. [Google Scholar] [CrossRef]
  84. Sun, K.; Wang, Y.; Chang, C.; Yang, S.; Di, S.; Niu, P.; Wang, S.; Li, L. Molten-salt synthesis of crystalline C3N4/C nanosheet with high sodium storage capability. Chem. Eng. J. 2021, 425, 131591. [Google Scholar] [CrossRef]
  85. Wang, J.; Wang, G.; Cheng, B.; Yu, J.; Fan, J. Sulfur-doped g-C3N4/TiO2 S-scheme heterojunction photocatalyst for Congo Red photodegradation. Chin. J. Catal. 2021, 42, 56–68. [Google Scholar] [CrossRef]
  86. Ghoreishian, S.M.; Raju, G.S.R.; Pavitra, E.; Kwak, C.H.; Han, Y.-K.; Huh, Y.S. Ultrasound-assisted heterogeneous degradation of tetracycline over flower-like rGO/CdWO4 hierarchical structures as robust solar-light-responsive photocatalysts: Optimization, kinetics, and mechanism. Appl. Surf. Sci. 2019, 489, 110–122. [Google Scholar] [CrossRef]
  87. Sun, F.; Xu, D.; Xie, Y.; Liu, F.; Wang, W.; Shao, H.; Ma, Q.; Yu, H.; Yu, W.; Dong, X. Tri-functional aerogel photocatalyst with an S-scheme heterojunction for the efficient removal of dyes and antibiotic and hydrogen generation. J. Colloid Interface Sci. 2022, 628, 614–626. [Google Scholar] [CrossRef] [PubMed]
  88. Luna, A.L.; Matter, F.; Schreck, M.; Wohlwend, J.; Tervoort, E.; Colbeau-Justin, C.; Niederberger, M. Monolithic metal-containing TiO2 aerogels assembled from crystalline pre-formed nanoparticles as efficient photocatalysts for H2 generation. Appl. Catal. B Environ. 2020, 267, 118660. [Google Scholar] [CrossRef]
  89. Sun, F.; Qi, H.; Xie, Y.; Xu, D.; Yu, W.; Ma, Q.; Yang, Y.; Yu, H.; Dong, X. Self-standing Janus nanofiber heterostructure photocatalyst with hydrogen production and degradation of methylene blue. J. Am. Ceram. Soc. 2022, 105, 1428–1441. [Google Scholar] [CrossRef]
  90. Pitcheri, R.; Mooni, S.P.; Harikrishnan, L.; Raghav, J.; Roy, S.; Maaouni, N.; Radhalayam, D.; Alothman, A.A.; Alharbi, A.F.; Al, Z.F.A.M.; et al. Novel S-scheme β-Cu2V2O7/Ni/Pg-C3N4 heterojunction photocatalyst for sunlight-induced degradation of RhB. Surf. Interfaces 2024, 52, 104950. [Google Scholar] [CrossRef]
  91. Madima, N.; Kefeni, K.K.; Mishra, S.B.; Mishra, A.K. TiO2-modified g-C3N4 nanocomposite for photocatalytic degradation of organic dyes in aqueous solution. Heliyon 2022, 8, e10683. [Google Scholar] [CrossRef] [PubMed]
  92. Khamesan, A.; Esfahani, M.M.; Ghasemi, J.B.; Farzin, F.; Parsaei-Khomami, A.; Mousavi, M. Graphitic-C3N4/ZnCr-layered double hydroxide 2D/2D nanosheet heterojunction: Mesoporous photocatalyst for advanced oxidation of azo dyes with in situ produced H2O2. Adv. Powder Technol. 2022, 33, 103777. [Google Scholar] [CrossRef]
  93. Xie, Q.; He, W.; Liu, S.; Li, C.; Zhang, J.; Wong, P.K. Bifunctional S-scheme g-C3N4/Bi/BiVO4 hybrid photocatalysts toward artificial carbon cycling. Chin. J. Catal. 2020, 41, 140–153. [Google Scholar] [CrossRef]
  94. Mallah, A.; Abdelrahman, E.A.; Raza, N.; Alqarni, L.S.; Ismail, M.; Modwi, A. Boosting of photocatalytic degradation of Malachite green dye on facile synthesized highly active Ca@TiO2@g-C3N4 photocatalyst: Photocatalytic mechanism and kinetics. Inorg. Chem. Commun. 2024, 170, 113464. [Google Scholar] [CrossRef]
  95. Madona, J.; Sridevi, C.; Indumathi, N.; Gokulavani, G.; Velraj, G. A novel carbon doped CeO2/g-C3N4 heterostructure for disinfection of microorganisms and degradation of Malachite green and Amoxicillin under sunlight. Surf. Interfaces 2024, 44, 103803. [Google Scholar] [CrossRef]
  96. Rana, A.; Sonu, S.; Sudhaik, A.; Kumar, R.; Chawla, A.; Raizada, P.; Chaudhary, V.; Ahamad, T.; Kaya, S.; Kumar, N.; et al. Tailoring dual S-Scheme based g-C3N4/ZnO/TiO2 ternary photocatalytic system immobilized on floating cork for environmental remediation. J. Taiwan Inst. Chem. Eng. 2025, 168, 105914. [Google Scholar] [CrossRef]
  97. Gou, L.; Wang, W.; Liu, E.-Z.; Xu, L.; He, R.; Yang, Y. Fabrication of MOF-derived CoTiO3/g-C3N4 S-scheme heterojunction for photocatalyst wastewater treatment. J. Alloys Compd. 2022, 918, 165698. [Google Scholar] [CrossRef]
  98. Onwudiwe, D.C.; Olatunde, O.C.; Nkwe, V.M.; Ben Smida, Y.; Ferjani, H. Dual S-scheme heterojunction g-C3N4/Bi2S3/CuS composite with enhanced photocatalytic activity for methyl orange degradation. Inorg. Chem. Commun. 2023, 155, 111075. [Google Scholar] [CrossRef]
  99. Duan, X.; Yang, J.; Zhu, J.; Li, H.; Fang, Y.; Liu, R.; Yang, C.; Liu, W.; Ding, C.; Liu, Q.; et al. Activated CdS/sulfur doped g-C3N4 photocatalyst for dye and antibiotic degradation: Experimental and DFT verification of S-scheme heterojunction. Environ. Res. 2025, 266, 120487. [Google Scholar] [CrossRef]
  100. Liu, H.; Sun, F.; Li, X.; Ma, Q.; Liu, G.; Yu, H.; Yu, W.; Dong, X.; Su, Z. g-C3N4/TiO2/ZnIn2S4 graphene aerogel photocatalysts with double S-scheme heterostructure for improving photocatalytic multifunctional performances. Compos. Part B Eng. 2023, 259, 110746. [Google Scholar] [CrossRef]
  101. Yang, W.; Tang, S.; Wei, Z.; Chen, X.; Ma, C.; Duan, J.; Tan, R. Separate-free BiPO4/graphene aerogel with 3D network structure for efficient photocatalytic mineralization by adsorption enrichment and photocatalytic degradation. Chem. Eng. J. 2021, 421, 129720. [Google Scholar] [CrossRef]
  102. Yao, Y.; Yang, Y.; Wang, Y.; Zhang, H.; Tang, H.; Zhang, H.; Zhang, G.; Wang, Y.; Zhang, F.; Yan, H. Photo-induced synthesis of ternary Pt/rGO/COF photocatalyst with Pt nanoparticles precisely anchored on rGO for efficient visible-light-driven H2 evolution. J. Colloid Interface Sci. 2022, 608, 2613–2622. [Google Scholar] [CrossRef] [PubMed]
  103. Yeh, T.; Cihlar, J.; Chang, C.; Cheng, C.; Teng, H. Roles of graphene oxide in photocatalytic water splitting. Mater. Today 2013, 16, 78–84. [Google Scholar] [CrossRef]
  104. Dong, S.; Cui, L.; Tian, Y.; Xia, L.; Wu, Y.; Yu, J.; Bagley, D.M.; Sun, J.; Fan, M. A novel and high-performance double Z-scheme photocatalyst ZnO-SnO2-Zn2SnO4 for effective removal of the biological toxicity of antibiotics. J. Hazard. Mater. 2020, 399, 123017. [Google Scholar] [CrossRef] [PubMed]
  105. Liu, H.; Li, X.; Ma, L.; Sun, F.; Yue, B.; Ma, Q.; Wang, J.; Liu, G.; Yu, H.; Yu, W.; et al. A three-dimensional TiO2/C/BiOBr graphene aerogel for enhancing photocatalysis and bidirectional sulfur conversion reactions in lithium-sulfur batteries. J. Environ. Chem. Eng. 2022, 10, 108606. [Google Scholar] [CrossRef]
  106. Jabbar, Z.H.; Okab, A.A.; Graimed, B.H.; Issa, M.A.; Ammar, S.H. Photocatalytic destruction of Congo red dye in wastewater using a novel Ag2WO4/Bi2S3 nanocomposite decorated g-C3N4 nanosheet as ternary S-scheme heterojunction: Improving the charge transfer efficiency. Diam. Relat. Mater. 2023, 133, 109711. [Google Scholar] [CrossRef]
  107. Liu, K.; Chen, J.; Sun, F.; Yu, J.; Zhang, X.; Xu, Y.; Liu, Y.; Tang, M.; Yang, Y. Enhanced degradation of azo dyes wastewater by S-scheme heterojunctions photocatalyst g-C3N4/MoS2 intimately coupled Rhodopseudomonas palustris with chitosan modified polyurethane sponge carrier. Int. J. Hydrogen Energy 2023, 48, 22319–22333. [Google Scholar] [CrossRef]
  108. Hassan, A.E.; Elsayed, M.H.; Hussien, M.S.A.; Mohamed, M.G.; Kuo, S.; Chou, H.; Yahia, I.S.; Mohamed, T.A.; Wen, Z. V2O5 nanoribbons/N-deficient g-C3N4 heterostructure for enhanced visible-light photocatalytic performance. Int. J. Hydrogen Energy 2023, 48, 9620–9635. [Google Scholar] [CrossRef]
  109. Yaqoubi, M.; Ghanbari, M.; Maya, R.W.; Jasim, H.A.; Salavati-Niasari, M. Synthesis and characterization of S-Scheme NiMn2O4/g-C3N4 nanocomposites heterojunction photocatalyst for effective degradation of organic pollutants under visible light. Results Eng. 2025, 25, 104037. [Google Scholar] [CrossRef]
  110. Alqarni, L.S.; Alghamdi, M.D.; Alhussain, H.; Elamin, N.Y.; Taha, K.K.; Modwi, A. S-scheme MgO-TiO2@g-C3N4 nanostructures as efficient photocatalyst for alizarin red S photodegradation. J. Mater. Sci. Mater. Electron. 2024, 35, 11996. [Google Scholar] [CrossRef]
  111. Hosseini, M.; Ghanbari, M.; Dawi, E.A.; Shuhata Alubiady, M.H.; Al-Ani, A.M.; Alkaim, A.F.; Salavati-Niasari, M. CaSnO3/g-C3N4 S-scheme heterojunction photocatalyst for the elimination of erythrosine and eriochrome black T from water under visible light. Results Eng. 2024, 21, 101903. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of g-C3N4-based S-scheme heterojunction photocatalytic degradation of dye.
Figure 1. Schematic diagram of g-C3N4-based S-scheme heterojunction photocatalytic degradation of dye.
Catalysts 15 00592 g001
Figure 2. (a) A schematic diagram of an RCN-VO S-scheme heterojunction for the photocatalytic degradation of tetracycline [18]. (b) A schematic diagram of the photocatalytic degradation of amoxicillin by a V2O5-g-C3N4 S-scheme nanocomposite material. (c) The preparation process of VO/CNNS composite materials [28]. (d) A schematic diagram of the GO/g-C3N4/TiO2 heterojunction photocatalyst and (e,f) HRTEM images of the GO/g-C3N4/TiO2 heterojunction photocatalyst; (e) is an enlarged image of the selected area in (f) (marked with a red border) [29].
Figure 2. (a) A schematic diagram of an RCN-VO S-scheme heterojunction for the photocatalytic degradation of tetracycline [18]. (b) A schematic diagram of the photocatalytic degradation of amoxicillin by a V2O5-g-C3N4 S-scheme nanocomposite material. (c) The preparation process of VO/CNNS composite materials [28]. (d) A schematic diagram of the GO/g-C3N4/TiO2 heterojunction photocatalyst and (e,f) HRTEM images of the GO/g-C3N4/TiO2 heterojunction photocatalyst; (e) is an enlarged image of the selected area in (f) (marked with a red border) [29].
Catalysts 15 00592 g002
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

Song, X.; Ma, Z.; Wang, Z.; Jin, S.; Hu, J.; Xu, P.; Chen, Y. Graphitic Carbon Nitride-Based S-Scheme Heterojunctions: Recent Advances in Photocatalytic Dye Degradation. Catalysts 2025, 15, 592. https://doi.org/10.3390/catal15060592

AMA Style

Song X, Ma Z, Wang Z, Jin S, Hu J, Xu P, Chen Y. Graphitic Carbon Nitride-Based S-Scheme Heterojunctions: Recent Advances in Photocatalytic Dye Degradation. Catalysts. 2025; 15(6):592. https://doi.org/10.3390/catal15060592

Chicago/Turabian Style

Song, Xiaofang, Zhenxing Ma, Zhiyong Wang, Shiyi Jin, Jingding Hu, Penghui Xu, and Yijiang Chen. 2025. "Graphitic Carbon Nitride-Based S-Scheme Heterojunctions: Recent Advances in Photocatalytic Dye Degradation" Catalysts 15, no. 6: 592. https://doi.org/10.3390/catal15060592

APA Style

Song, X., Ma, Z., Wang, Z., Jin, S., Hu, J., Xu, P., & Chen, Y. (2025). Graphitic Carbon Nitride-Based S-Scheme Heterojunctions: Recent Advances in Photocatalytic Dye Degradation. Catalysts, 15(6), 592. https://doi.org/10.3390/catal15060592

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