Next Article in Journal
Conjugate Heat Transfer Analysis for Cooling of a Conductive Panel by Combined Utilization of Nanoimpinging Jets and Double Rotating Cylinders
Next Article in Special Issue
Redox-Enhanced Photoelectrochemical Activity in PHV/CdS Hybrid Film
Previous Article in Journal
Chitosan-Graphene Oxide Dip-Coated Polyacrylonitrile-Ethylenediamine Electrospun Nanofiber Membrane for Removal of the Dye Stuffs Methylene Blue and Congo Red
Previous Article in Special Issue
Pt-Based Oxygen Reduction Reaction Catalysts in Proton Exchange Membrane Fuel Cells: Controllable Preparation and Structural Design of Catalytic Layer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Engineered g-C3N5-Based Nanomaterials for Photocatalytic Energy Conversion and Environmental Remediation

1
State Key Laboratory of Petroleum Pollution Control, China University of Petroleum (East China), Qingdao 266580, China
2
Shandong Engineering and Technology Research Center for Ecological Fragile Belt of Yellow River Delta, Binzhou University, Binzhou 256600, China
3
School of Energy and Power Engineering, Jiangsu University, Zhenjiang 212013, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(3), 499; https://doi.org/10.3390/nano13030499
Submission received: 15 December 2022 / Revised: 9 January 2023 / Accepted: 23 January 2023 / Published: 26 January 2023
(This article belongs to the Special Issue Nanomaterials for Photochemical/Photoelectrochemical Application)

Abstract

:
Photocatalysis plays a vital role in sustainable energy conversion and environmental remediation because of its economic, eco-friendly, and effective characteristics. Nitrogen-rich graphitic carbon nitride (g-C3N5) has received worldwide interest owing to its facile accessibility, metal-free nature, and appealing electronic band structure. This review summarizes the latest progress for g-C3N5-based photocatalysts in energy and environmental applications. It begins with the synthesis of pristine g-C3N5 materials with various topologies, followed by several engineering strategies for g-C3N5, such as elemental doping, defect engineering, and heterojunction creation. In addition, the applications in energy conversion (H2 evolution, CO2 reduction, and N2 fixation) and environmental remediation (NO purification and aqueous pollutant degradation) are discussed. Finally, a summary and some inspiring perspectives on the challenges and possibilities of g-C3N5-based materials are presented. It is believed that this review will promote the development of emerging g-C3N5-based photocatalysts for more efficiency in energy conversion and environmental remediation.

1. Introduction

As civilization has developed, energy crises and environmental contaminations have become major obstacles to the further development of human society [1,2,3,4]. Semiconductor-based photocatalysis is regarded as one of the most efficient technologies to address serious energy and environmental issues [5,6,7]. The mass production of highly efficient, dependable, and reasonably priced photocatalysts with strong charge carrier segregation, plenty of active sites, a broad optical absorption range, and high redox potentials is crucial for the commercialization of photocatalytic technology [8,9,10]. As a result, much effort has been devoted to designing more robust and efficient photocatalysts.
Graphitic carbon nitride (g-C3N4) has sparked extensive interest because of its metal-free nature, facile accessibility, excellent physicochemical stability, and appealing electronic band structure [5,6,11]. However, the inherent disadvantage of carrier recombination and the restricted light absorption range dramatically hampered the practical application of g-C3N4 [12,13]. To elevate the photocatalytic performance of g-C3N4, versatile modification procedures, such as doping, heterojunctions, and chemical structural regulation (tuning the C/N ratio), have been implemented [14,15]. Compared with g-C3N4, g-C3N5 has piqued particular interest because of its nitrogen-rich moiety and sp2 hybridized atoms, which assist in the optimization of the electronic band structure [16].
The researchers have recently synthesized 2D g-C3N5 with different structures using diverse precursors. For example, g-C3N5 with heptazine moieties linked together by azo linkage (−N=N−) was synthesized by thermal deamination of 2,5,8-trihydrazino-s-heptazine [17]. Terminal triazole-based g-C3N5 was fabricated via straightforward self-condensation of 3-amino-1,2,4-triazole [18,19,20], while g-C3N5 with one-triazole and two-triazine combinations was formed via self-assembly of 5-amino-1H-tetrazole (5-ATTZ) [21]. Furthermore, a variety of techniques for modifying g-C3N5 have been developed, including element doping, defect engineering, and heterojunction engineering. To date, g-C3N5-based materials have been widely employed in energy conversion and environmental remediation as a new style of carbon nitride specimen [17,22,23,24]. However, a comprehensive review of g-C3N5-based photocatalysts for energy and environmental applications is still lacking.
This review summarizes the latest progress of g-C3N5-based photocatalysts for energy and environmental applications. Various synthesis topologies of pristine g-C3N5 and several engineering strategies for g-C3N5 (such as elemental doping, defect engineering, and heterojunction creation) are presented. In addition, the photocatalytic applications in H2 evolution, CO2 reduction, N2 fixation, NO purification, and aqueous pollutant degradation are discussed. To the best of our knowledge, this is the first review of g-C3N5-based photocatalysts. It is anticipated that this review will provide readers with a deep understanding of emerging g-C3N5-based photocatalysts, which might promote the design of more robust metal-free photocatalysts.

2. Synthesis Strategies of Pristine g-C3N5

In this section, the synthesis techniques and mechanisms of pristine g-C3N5 with various structures derived from various precursors will be thoroughly explored. The pristine g-C3N5 can be divided into three types, i.e., g-C3N5 with triazine and triazole (Figure 1a), g-C3N5 with terminal triazole (Figure 1b), and g-C3N5 with azo-linked heptazine (Figure 1c). They can be synthesized by one-step polymerization of inexpensive nitrogen-rich precursors such as 5-amino-1H-tetrazole [21], 3-amino-1,2,4-triazole [18,19,25], and melamine [17].

2.1. g-C3N5 Containing One-Triazole and Two-Triazine

Unlike g-C3N4, which has three triazine moieties, g-C3N5 has one triazole and two triazine moieties, with one triazole ring replacing one triazine ring in g-C3N4. Compared with triazine or heptazine moieties, the triazole moiety possesses more electrons and more pyrrolic N sites than triazine or heptazine, which provides more numbers of the basic sites and enhances the basic catalytic activity. In this section, we will describe the templated and template-free methods for such g-C3N5 preparations.

2.1.1. The Hard Template Approach

According to the literature, KIT-6 and SBA-15 are two commonly used templates for the synthesis of g-C3N5. Vinu and co-workers prepared porous g-C3N5 (MCN-11) via a simple self-assembly of the 5-amino-1H-tetrazole (5-ATTZ) in the presence of hard template KIT-6. In brief, 5-ATTZ was impregnated into the pore channels of the KIT-6 template, carbonized for 4 h at 400 °C, and the template was removed with 5% HF acid to obtain MCN-11. It was demonstrated that the structure of MCN-11 was composed of one triazole and two triazine moieties by DFT calculations (Figure 2a–d) and spectroscopic analyses (Figure 2e,f) [21]. Additionally, the impact of temperature on the material’s nitrogen content, structure, and surface functional groups was also investigated. As the carbonization temperature ascended from 350 to 550 °C, the C/N ratio increased from 1.43 to 1.62, representing the structural change from g-C3N5 to g-C3N4. g-C3N5 containing 1-NH2/NH-1,2,4-triazole units (MCN-11) (Figure 2g) was obtained at 350 and 400 °C, whereas g-C3N4 containing 2-NH2/NH-1,3,5-triazine units (Figure 2h) was obtained at 450, 500, and 550 °C [26]. Vinu’s group also synthesized g-C3N5 nanorods with a combination of triazine and triazole groups, named MCN-14-X, using SBA-15 as a hard template instead of KIT-6. The preparation process is briefly summarized as follows: 5-ATTZ was impregnated into the porosity canals of the SBA-15 template, then carbonized for 4 h at 400 °C, and the template was etched with 5% HF acid to obtain MCN-14 [27]. The obtained mesoporous carbon nitrides, MCN-11 and MCN-14, have a good effect on the adsorption and conversion of CO2 and are the catalysts for Knoevenagel condensation. However, this hard template still suffers from cumbersome template preparation processes and an unfriendly template removal process.

2.1.2. The Template-Free Approach

A template-free approach was reported by Zhang et al. [28]. Two-dimension metal-free g-C3N5 nanosheets were acquired by the polymerization of 3-AT at 500℃ for 3 h under a half-cover air using the heating rate of 5 °C min−1. The prepared g-C3N5 exhibits excellent crystal properties, which can be confirmed by the peaks of 100 (in-plane structural ordering) and 002 (interlayer structural ordering) in XRD characterization (Figure 3a). XPS spectra (Figure 3b–d) confirmed the presence of one-triazole and two-triazine. Compared with g-C3N4, g-C3N5 exhibits a better visible light response (Figure 3e), a lower band gap (Figure 3f), and a more effective separation of photogenerated electrons and holes (Figure 3g). It may be due to the abundance of unpaired electrons on g-C3N5, confirmed by the ESR spectra (Figure 3h). Meanwhile, the orbital calculations (Figure 3i,j) also confirm that g-C3N5 has better photo-generated carrier separation efficiency than g-C3N4. Therefore, the prepared g-C3N5 displays better photocatalytic performance than g-C3N4. This approach is simple, time-saving, energy-saving, and environment-friendly, which makes it conducive to large-scale production and application of materials.

2.2. g-C3N5 with Terminal Triazole

The structure of g-C3N5 with terminal triazoles is constructed by 1,2,4-triazoles and 1H-1,2,3-triazoles, which replace two hydrogens on the terminal amino moiety of one triazine ring in g-C3N4. This replacement both enhanced conjugation and increased nitrogen content (-NH2, -NH), which was beneficial to expand the photo-absorption range and enhancing the selectivity and adsorption of acidic gases, respectively. In this section, we will summarize the methods for producing terminal triazole-based g-C3N5 using KIT-6 and KBr templates.

2.2.1. The Method Using the KIT-6 Template

Vinu et al. [18] have achieved a remarkable breakthrough in the synthesis of mesoporous g-C3N5, which was named MCN-8. It was synthesized by a thermal condensation process of the 3-amino-1,2,4-triazole precursor in a KIT-6 silica template using a nano-casting technique. In order to detect the structures and properties, advanced characterization techniques are used. XRD measurements reveal the structural ordering and graphitic character of MCN-8 (Figure 4a). According to the low-angle powder XRD pattern, the MCN-8 materials have a highly organized three-dimensional porous mesostructure. The diffraction peak at 2θ = 27.28 (002) verifies the CN framework’s strong crystallinity and interlayer turbostratic ordering, as does a faint peak at 2θ = 13.48 (100), which confirms the in-plane structural ordering. The TEM (Figure 4d) and SEM (Figure 4e) images clearly demonstrate MCN-8’s lattice stripes and porous structure. The N2 adsorption-desorption measurement (Figure 4b) also reveals that the MCN-8 has organized mesopores and a large specific surface area, both of which contribute to improved adsorption. The near-edge X-ray absorption fine structure (NEXAFS) (Figure 4c), high-resolution XPS (Figure 4f), and energy loss (EEL) spectra (Figure 4g) demonstrate the existence of the terminal triazole unit and tri-s-triazine ring in the CN framework. This chemical structure not only reduces the bandgap to 2.2 eV (Figure 4h), which is crucial for extending visible light absorption for a faster rate of H2 evolution (Figure 4i), but it also demonstrates improved selective sensing of toxic acid molecules (Figure 4j). However, the fabrication of the nano-template (KIT-6) requires a high temperature (540 °C) and a long time, which consumes both energy and time. In order to save energy, Vinu and co-workers [19] used ethanol washing instead of high-temperature calcination to obtain the KIT-6 for synthesizing 3D mesoporous g-C3N5 (MCN-8E). It was found that the porous structure and graphite essence of MCN-8E are strikingly similar to those of MCN-8. Considering the use of a hazardous etching chemical (HF) for template removal causes the loss of active sites and degrades photocatalytic efficacy as well as the cumbersome preparation of the template KIT-6, the search for an energy-saving, non-toxic, and environment-friendly synthesis route has attracted the attention of researchers.

2.2.2. The Method Using the KBr Template

Greener synthesis approaches are urgently needed to eliminate the use of the toxic HF etchant in g-C3N5 production. Wang et al. [20] fabricated g-C3N5 with mesoporous and rod-like morphology (RN-g-C3N5) via thermal condensation of the 3-amino-1,2,4-triazole with KBr as the template. Figure 5a,b reveal that the prepared RN-g-C3N5 has a rod-like porous structure due to the guiding effect of KBr. This special structure expands the visible adsorption range (Figure 5c), reduces the bandgap to 1.90 eV (Figure 5d), and improves the separation of electron-hole pairs (Figure 5e,f). As a result, the RN-g-C3N5 displays superior photocatalytic performance and reusability in MB degradation (Figure 5g,h). It shows greater potential for applications in the field of environmental remediation. Compared with the traditional silica templates, the KBr template is an environmentally friendly strategy for avoiding structural damage, the loss of active sites, and significant environmental contamination caused by HF. In short, this procedure is straightforward, environmentally friendly, and inexpensive.

2.3. g-C3N5 Consisted of Heptazine Units Bridged by Azo

g-C3N5, consisting of heptazine units (tri−s−triazine) bridged by azo, is an analogue of g-C3N4, which is derived by replacing the tertiary nitrogen in the g-C3N4 structure with azo. The extended π−conjugation formed by coupling the orbitals of the azo-linkage with the π-conjugated structures of the heptazine narrowed its bandgap, which expanded the optical absorption range. The theoretical design and experimental fabrication of g-C3N5 (−N = N−) will be summarized in this section.

2.3.1. Theoretical Design

Cao et al. [29] designed a 2D crystalline graphitic g-C3N5 material by employing the heptazine structure as the node and the azo group as the linker. Density functional theory (DFT) was employed to investigate the material’s optical, electrical, and catalytic characteristics. It was found that 2D g-C3N5 has an excellent performance in non-metal oxygen reduction reactions, gas adsorption, separation, and conversion.

2.3.2. Experimental Fabrication

In terms of experimental fabrications, two methods for synthesizing the g-C3N5 framework are presented. As shown in Figure 6a, Kumar et al. [17] synthesized this g-C3N5 via thermal deamination of melem hydrazine (MH), also named 2,5,8-trihydrazino-s-heptazine, which was obtained by polymerization of melamine at 425 °C for overnight, followed by a hydrothermal process with hydrazine hydrate. Characterization techniques such as XPS (Figure 6c) and NMR (Figure 6d) indicated that the structure is made up of two s-heptazine units connected by an azo linkage. The overlap of azo atoms with the aromatic network of heptazine units extends conjugation and increases electron density on the heptazine nucleus (Figure 6b). The lower EPR signal strength of g-C3N5 compared to g-C3N4 indicates that the number of lone pair electrons in g-C3N5 drops, confirming that extra N atoms bond with tertiary nitrogen atoms outside the heptazine nucleus to form azo rather than substitute C atoms in the heptazine motif (Figure 6h). This structure increases visible light responsiveness and photogenerated electron-hole separation, as seen by increased visible light absorption in Figure 6e, lower PL intensity in Figure 6f, and a shorter average lifetime of the photogenerated charge carrier in Figure 6g. The 2D g-C3N5 containing s-heptazine units and azo was successfully prepared by this approach, although the process was complicated and time-consuming. In another work, Liu et al. synthesized g-C3N5 nanosheets by heating a mixture of NH4Cl and 3-amino-1,2,4-triazole (3-AT), followed by protonation exfoliation. It was found that the morphology of the nanosheet increases the specific surface area, facilitates charge separation, and modifies the band structure [30].
In summary, the preparation methods of g-C3N5 with three different structures, namely, g-C3N5 with a terminal triazole, g-C3N5 containing s-heptazine linked by azo, and g-C3N5 containing 1 triazole and 2 triazines, are presented. It has been witnessed that breakthroughs and progress have been made in the field of g-C3N5 synthesis, from scratch and complexity to simplicity. It is necessary to further explore more environmentally friendly, energy-saving, time-saving, and labor-saving preparation methods.

3. Functional Engineering of g-C3N5

Pristine g-C3N5 performs better in the applications of CO2 adsorption and conversion, pollutant degradation, and H2 evolution. However, pristine g-C3N5 still suffers electron-hole recombination and limited use of visible light. To address the aforementioned issues, g-C3N5 functional materials have been designed and fabricated to broaden their uses in a variety of domains. We will go over the preparations of g-C3N5 functionalized materials in this section, including defect engineering and heterojunction engineering.

3.1. Defect Engineering

One strategy for improving the photocatalytic properties of g-C3N5 is to introduce structural defects into the surface. Defects are produced by the dislocation of an atom, which breaks the periodic arrangement. Doping and vacancy are the two types of engineering defects. Doping usually involves adding atoms to the original structure, while vacancies occur when a host atom is losing its position in the original crystal structure. The defects can cause electronic redistribution in the material, which can regulate spectral absorption and affect the excitation, migration, and recombination of photogenerated electron holes.

3.1.1. Doping Design

Doping is a good strategy because it not only changes the electronic structure, resulting in the band gap and the visible light absorption variation, but it also changes the surface properties, which affect the photocatalytic performance. Next, we will discuss metal doping, non-metal doping, and both metal and non-metal doping in the functional engineering of g-C3N5.
Metal doping can create defect states in the band structure of g-C3N5, which trap holes or electrons to change interface charge transfer, inhibiting the recombination of photogenerated holes and electrons. As a new type of N-rich material, the N-rich sites of g-C3N5 with stronger electronegativity are easy to combine with positive metal ions. Transition metals with unoccupied d orbitals, which may function as electron donors or acceptors, can easily form complexes with g-C3N5 through electron transfer. The feasibility of transition metal doping g-C3N5 is verified by Niu’s team [31] through first-principles calculations. They designed a set of models for depositing transition metal (TM) atoms (V, Cr, Mn, Fe, Co, Ni, Cu, Mo, Ru, Pd, W, and Pt) on g-C3N5 to construct transition metal (TM) doped g-C3N5 single-atom catalysts (SACs), TM-g-C3N5. The adsorption energies calculated by DFT are all negative, which demonstrates that the structures of all 12 TM-g-C3N5 are stable. This study has significant implications for future research into g-C3N5-based stable photocatalytic materials.
In terms of material fabrication, Fang et al. [32] prepared pristine g-C3N5 according to Zhang’s method [28] and pyrolyzed a mixture of it with potassium chloride, lithium chloride, and cobalt chloride dihydrate to prepare cobalt-doped g-C3N5, whose doping pattern is shown in Figure 7a. HR-TEM images (Figure 7b), HAADF-STEM images (Figure 7c), and elemental mapping (Figure 7d–f) all show that individual nitrogen atoms are uniformly doped into g-C3N5. In addition, it was also confirmed that the configuration of Co-N4 was formed by the combination of Co with pyridine nitrogen in g-C3N5, as confirmed by the XANES spectra (Figure 7g) and R-space of XANES (Figure 7h). The prepared Co-C3N5 has stable catalytic performance for PCB28 degradation, which is attributed to sulfate radicals from PMS activated by Co-N4. Similar work has also been done by Gan et al. [33]. They prepared pristine g-C3N5 according to Vinu’s method [18] and pyrolyzed the mixture of pristine g-C3N5 with 1,10-phenanthroline monohydrate and (CH₃COO)₂Co·4H₂O to prepare another cobalt-doped g-C3N5. The structure of Co-N binding was confirmed by advanced characterization methods, such as a peak at 2170 cm−1 in FT-IR spectra (Figure 7i) and a Co-Nx peak in the XPS spectra of Co2p (Figure 7j). Cobalt doping significantly enhances optical properties, such as reducing the band gap from 1.80 to 1.63 eV in Tauc’s plot (Figure 7k), which helps to activate PMS for the generation of active species, including free radicals (·O2, SO4·, and ·OH) and non-free radicals (1O2 and tetravalent cobalt oxide). In SMX degradation experiments, it can be degraded with a removal rate of 99.57% in 20 min. The tetravalent cobalt oxide performed a vital role, whereas the other species made a minor contribution. In addition, Vinu’s team [34] synthesized mesoporous titanium carbonitride, named MTiCN. It was obtained by carbonizing the mixture of mesoporous C3N5 and titanium tetrachloride at 900 °C for 5 h. The resulting MTiCN samples have a rod-like morphology, a mesoporous structure, and a large specific area, as well as a high carbon content. It shows excellent H2 evolution performance similar to commercial platinum/carbon. Moreover, Vivek Polshettiwar [35] synthesized K-doped g-C3N5 (K-g-C3N5) via the thermal polymerization of 3-AT and KBr. A dipole interaction is generated between K and N atoms as opposed to K atoms replacing N atoms, hence altering the distribution of electrons and narrowing the band gap to increase light absorption. As an electron trap, K also facilitates the transfer and separation of photogenerated carriers.
Non-metallic doping is another effective method of modifying g-C3N5. Since it has a high electron affinity and electronegativity, it may easily substitute the N atom in g-C3N5 and establish covalent bonds with the C atom. Non-metallic ions penetrate the lattice and act as impurity energy levels in the valence band, therefore reducing the band gap, separating electrons and holes, and enhancing the catalytic properties. To date, non-metallic components such as N, B, P, etc. have been employed to modify g-C3N5.
Hu et al. [36] investigated the phosphorus doping of g-C3N5 (P-g-C3N5). P-doped triazole-based g-C3N5 was produced by thermal condensation at 500 °C for 4 h, followed by thermal copolymerization at 500 °C for another 4 h, using 3-AT and HCCP as precursors. Figure 8a depicts its doping structures: P−N/P = N structures were formed between P and N, which were derived from the XAFS spectra (Figure 8b) and EXAFS spectra (Figure 8c). P2P orbitals provide donor levels in the forbidden band to donate more electrons, narrowing the band gap and increasing the visible light usage rate, as confirmed by UV-vis spectra (Figure 8d). Compared with the undoped g-C3N5 specimen, P-g-C3N5 demonstrated a higher separation efficiency of photogenerated charges, which was confirmed by PL spectra (Figure 8e). As a result, the P-g-C3N5 displays an excellent rhodamine B and tetracycline degradation rate.
Li et al. [38] constructed 18 varieties of B-doped g-C3N5, including substitution doping and pore doping, followed by investigating the impact of B-doping on the material’s photo-absorption range, band structure, and charge separation. The results revealed that the optimal impurity levels increased light absorption, minimized the recombination of photogenerated charge, and improved the performance of the catalyst. In terms of material preparation, Qu et al. [37] synthesized boron-doped g-C3N5 (B-C3N5) by a one-step pyrolysis process employing ammonium borate and 3-AT, whose structure is shown in Figure 8f. The increased ID/IG (Figure 8h) shows that the graphite-like structure was distorted, confirming the B as the heteroatom doped into the structure of the CN. Indeed, the decreased C/N ratio calculated by element analysis confirmed the loss of carbon atoms. The doping of B makes up for the loss of the C atom, as confirmed by the decreased ESR spectra intensity of B-C3N5 (Figure 8g). That is to say, B tends to occupy the C vacancy sites in C3N5 to form the B-O-H functional group, which can expand the light absorption range (Figure 8i), enhance the charge transfer ability (Figure 8j), and optimize the structural stability. In the photocatalytic nitrogen fixation experiments, it can achieve a photocatalytic nitrogen fixation rate of 421.18 μmol·h−1·g−1, demonstrating its superior performance. The B-O-H groups play the most important role; B can absorb and activate nitrogen using its Lewis acid characteristics, and O-H can provide H protonate to activated nitrogen for NH3 formation using its Bronsted acidic characteristics.
Ding et al. [39] synthesized an N-doped carbon catalyst (PDA-g-CN) by pyrolyzing g-C3N5@PDA hybrid, which was created by the polymerization reaction of dopamine hydrochloride and g-C3N5 under mild alkaline conditions, at 800 °C for 2 h. The developed PDA-g-CN catalyst has shown good degradation efficiency for organic pollutants (SMX) via the non-radical mechanism. Nitrogen with a higher electronegativity makes the surface potential of carbon more positive, thereby enhancing the adsorption capacity of PMS and promoting the formation of high-redox potential C-PMS* complexes, which can oxidize organic contaminants (SMX) via electron transfer from SMX to C-PMS*.
Metal and nonmetal doping play a significant role in the modification of g-C3N5, which can alter the band gap, light absorption, and photogenerated charge separation. Most metal elements may enter the lattice by pore doping, whereas most non-metal elements may enter via substitution doping. On top of that, Ao et al. [40] synthesized the metal K and nonmetal I co-doped g-C3N5 (K, I-g-C3N5) via pyrolyzing 3-AT and KI. K may enter the lattice by pore doping, and I may enter via substitution doping. The band gap of K and I co-doped g-C3N5 is larger than that of pristine g-C3N5, which is unfavorable for light absorption. Fortunately, the more negative conduction band potential has a strong reduction ability and can reduce O2 to produce H2O2 (O2 + 2e + 2H+ → H2O2). It was also confirmed that the co-doping of metal and non-metal is an effective way to improve the properties of g-C3N5.

3.1.2. Vacancy Engineering

Vacancy defects also play an important role in enhancing photocatalytic capacity, which can be summarized as having three important aspects: light absorption ability, carrier separation efficiency, and surface reaction ability. In the field of materials, nitrogen vacancies and oxygen vacancies are commonly used to regulate photocatalytic performance. For example, Wang et al. [25] fabricated a triazole-based g-C3N5 material with N vacancies (Nv-g-C3N5) through the thermal polymerization of 3-AT assisted by NaOH. Later, Zhang et al. [41] used this method to fabricate the same Nv-g-C3N5 for constructing the NV-g-C3N5/BiOBr heterojunction. The structure is shown in Figure 9a. A strong Lorentz signal peak observed in the ESR curve of Nv-g-C3N5 confirmed the existence of nitrogen vacancies (Figure 9c). The peak at 2170 cm−1 observed in the FT-IR spectrum suggests the formation of cyano groups (Figure 9b). The cyano-groups extend the photo-absorption range (Figure 9d), reduce the band gaps from 2.08 to 1.5 eV (Figure 9e), and use N vacancies as electron traps to capture and store excited electrons, which facilitate the separation of photogenerated charge carriers. The Nyquist radius of NV-g-C3N5 decreases with the increase of N vacancies in Figure 9f, which is conducive to electron and hole migration and separation. In addition, the PL intensity of Nv-g-C3N5 decreases with the increase of N vacancies in Figure 9g, also reflecting the enhanced separation rate of charge carriers. Both nitrogen vacancies and cyano groups affect the photocatalytic and photoelectrochemical properties of materials. We believe that this simple, effective, and low-cost vacancy engineering is beneficial to the wide application of g-C3N5 in energy conversion and environmental remediation.

3.2. g-C3N5-Based Heterojunctions

Constructing heterojunctions between the two semiconducting materials is an efficient strategy for facilitating charge transfer. The photocatalytic activity of g-C3N5 materials can be boosted by constructing heterojunctions with another semiconductor. Until now, many g-C3N5-based heterostructures have been synthesized for energy conversions and environmental governance, such as g-C3N5/BiOBr [41], Bi4O5Br2/g-C3N5 [42], Bi4O5I2/g-C3N5 [43], Bi2WO6/g-C3N5 [44], Bi2MoO6/g-C3N5 [45], CeTi2O6/g-C3N5 [46], Ag3PO4/g-C3N5 [47], Ag2CO3/g-C3N5 [48], AgCl/g-C3N5 [49], FeOCl/g-C3N5 [50], LaCoO3/g-C3N5 [51], and MOFs/g-C3N5 [52]. This section will give a brief overview of heterojunctions based on g-C3N5, including type-I heterojunction, type-II heterojunction, Z-scheme heterojunction or S-scheme heterojunction, and Schottky junction. Their photoinduced carrier transfer routes are shown in Figure 10.

3.2.1. Type-I Heterojunction

For a type-I heterojunction (Figure 10a), the energy band structure is nested, with the bottom of the conduction band and the top of the valence band of the narrow band material situated inside the forbidden band of the wide band material. Both holes and electrons are transmitted from semiconductor 1 to semiconductor 2, where they may combine. Consequently, this heterojunction restricts the separation of electrons and holes. There are few materials about type-I heterojunction among the g-C3N5-based heterojunctions. It is worth noting that the azo and heptazine rings expand the conjugation of g-C3N5, increasing the migration properties and separation efficiency of electrons and holes. Thus, g-C3N5 can be used to improve another semiconductor property via type-I heterojunction. For example, Alam et al. [53] fabricated CdS/C3N5 heterojunctions with a core-shell structure by wrapping CdS nanowires with g-C3N5 shells obtained via melamine thermal deamination. It greatly outperformed CdS nanorods alone in the removal of 4-nitrophenol and rhodamine B (RhB) (Figure 11a), which was attributed to the passivating effect generated by the g-C3N5 shell for CdS and the effective charge separation produced by the type-I heterojunctions between CdS and g-C3N5. Through type-I heterojunction, the electrons transfer from CdS to g-C3N5, and the holes also transfer from CdS to g-C3N5, thus promoting hole extraction on CdS. Furthermore, photoinduced electron migration and separation are facilitated by the expanded conjugation in g-C3N5. In addition, Wang et al. [54] fabricated another kind of CdS/C3N5 for enhanced H2 production. It is fabricated by the hydrothermal reaction of the uniform suspension solution obtained by adding g-C3N5 to an aqueous solution of NH2CSNH2 and Cd(NO3)2·4H2O at 180 °C for 12 h. The synthesized CdS/C3N5 greatly decreased the electron-hole recombination rate via type-I heterojunction, which can produce hydrogen more than four times faster than pristine g-C3N5 (Figure 11b). Furthermore, Wang et al. [55] successfully fabricated the g-C3N4/g-C3N5 VDWs junction via a facile calcination technique with the precursors 3-AT for g-C3N4 and melamine for g-C3N5. It exhibited outstanding RhB and TC-HCl degradation performance, which was attributable to the charge transfer from g-C3N4 nanolayers to g-C3N5 nanorods through a type-I heterojunction interface and then effectively activating molecular oxygen to generate ROS (Figure 11c).

3.2.2. Type-II Heterojunction

Type-II heterojunction (Figure 10b) has a staggered arrangement of energy bands; the CB and VB potentials of semiconductor 1 are higher than those of semiconductor 2. Under visible light irradiation, photoinduced electrons may migrate from the CB of semiconductor 1 to that of semiconductor 2, while photoinduced holes are transferred in the opposite direction, from semiconductor 2 to semiconductor 1. Photoinduced electrons and holes accumulated on semiconductors 2 and 1, respectively. Consequently, type-II heterojunctions seem to achieve charge carrier separations.
In this section, g-C3N5-based type-II heterojunction materials and their synthesis methods are briefly reviewed. The basic process is to prepare g-C3N5 material, then mix it with other materials through precipitation or hydrothermal reaction to obtain the corresponding heterojunction materials. For example, Sun et al. [42] constructed a g-C3N5/Bi4O5Br2 type-II heterojunction by in-situ growth of Bi4O5Br2 on g-C3N5 nanosheets through the hydrothermal method (Figure 12a). The interlayer morphology and N-rich structure enhanced visible light utilization, the bismuth-rich property of Bi4O5Br2 increased the excited charge carrier lifetime, and the type-II surface heterojunction (g-C3N5/Bi4O5Br2) promoted charge transfer and separation. As shown in Figure 12c, the reactive species ·O2, produced by the reaction between separated electrons on the CB of Bi4O5Br2 and the adsorbed O2, together with the separated holes on the VB of g-C3N5, improve the degradation efficiency of sulfathiazole (Figure 12b). Indeed, Meng et al. [56] constructed a 2D/2D C3N5/GO type-II heterojunction using a sonochemical self-assembly method, achieving a U(VI) removal rate of 96.1% (Figure 12d,e). The photoreduction ability of the electrons on the GO conduction band and the physical adsorption ability of GO itself play an important role (Figure 12f). In addition, p-n heterojunction, a special type-II heterojunction, is composed of p-type and n-type semiconductors. Under visible light irradiation, the photoexcited electrons in the conduction band of the p-type semiconductor move to the n-type semiconductor, driven by the built-in electric field, while the holes move in the opposite direction, forming a p-n heterojunction. For example, Zhang et al. [41] used a hydrothermal process to make NV-g-C3N5/BiOBr. Nitrogen vacancies and P-N heterojunction promote N2 adsorption and activation.
However, the theory of type-II heterojunctions has also been questioned. Yu et al. [57,58] have presented the inadequacies of the theory of type-II heterojunction from thermodynamic, kinetic, and energy viewpoints. First, the carrier migration to lower CB or higher VB reduces the redox ability of the material accompanying the energy loss. Secondly, this transfer is also kinetically unfavorable because of the repulsive forces between electrons and electrons and holes and holes.
Z-scheme or S-scheme heterojunction (Figure 10c) has the same interleaved energy band structure as a Type-II heterojunction but a completely different mechanism for charge transfer. Electrons migrate from a high Fermi energy level to a low Fermi energy level when two semiconductors come into contact, creating an internal electric field (IEF) and energy band edge bending. Additionally, there are also Coulomb interactions between electrons and holes. As a result of the three aforementioned actions, electrons and holes with low redox potential merge, leaving electrons and holes with high redox potential in the conduction band of semiconducting 1 and the valence band of semiconductor 2. As a result, Z- or S-scheme heterojunctions can achieve excellent separation of photogenerated carriers while retaining a high redox capacity.
Z-scheme heterojunction is developed on the basis of liquid-phase and solid-state Z-scheme heterojunction. Yu et al. [57,58] demonstrated that the charge transfer hypothesis of liquid-phase and all-solid Z-scheme heterostructures is unfavorable in thermodynamics and dynamics, proposing that the “Z-scheme” name be abandoned and substituted with a new name: S-scheme heterostructure. In accordance with the literature, we will discuss the S- and Z-scheme heterostructures, respectively.

3.2.3. Z-Scheme Photocatalyst

Z-scheme heterojunction (Figure 10c) has been shown to be a promising technique for improving a single semiconductor’s photocatalytic efficiency. It can also be employed to modify g-C3N5 materials without doubt. To enhance the photocatalytic performance of g-C3N5, C3N5-based Z-scheme heterojunctions composed of g-C3N5 and the other semiconductor have been investigated. Herein, g-C3N5-based Z-scheme heterojunction materials and their synthesis methods are briefly reviewed. The basic process is to prepare the g-C3N5 material first, then mix it with other materials through liquid phase mixing, in-situ precipitation, hydrothermal reaction, and calcination to obtain the corresponding heterojunction materials. For example, Shi et al. [52] fabricated NH2-UiO-66/N-CN Z-scheme heterojunction photocatalysts via assembly in liquid mixing (Figure 13a). Taking Pt as a cocatalyst, the fabricated NH2-UiO-66/N-CN-2 displays a high hydrogen evolution rate (Figure 13b), which is attributed to the separation of photo-generated carriers facilitated by the close contact interface of Z-scheme heterojunction. This charge transfer mechanism was also verified by the density functional theory calculation (Figure 13c). This provides an effective method to improve the performance of the g-C3N5 photocatalyst by adjusting the electron distribution to promote charge separation through heterojunction interface engineering.

3.2.4. S-Scheme Heterojunction

The reported C3N5-based S-scheme heterostructure consisting of g-C3N5 and another oxidizing semiconductor may improve the photocatalytic activity of g-C3N5. Through charge transfer at the interface of the S-scheme heterojunction, the holes in g-C3N5 may be coupled with the electrons in another semiconductor through charge transfer at the interface of an S-scheme heterojunction. The electrons in g-C3N5 and holes in another semiconductor were preserved, which successfully separates electrons and holes and maintains a high redox capability. Herein, g-C3N5-based S-scheme heterojunction materials and their synthesis methods are briefly reviewed. The basic process is to prepare the g-C3N5 material first, then mix it with other materials through wet chemical synthesis, hydrothermal reactions, and a solvothermal approach to obtain the corresponding heterojunction materials. For example, Li et al. [45] constructed a Bi2MoO6/C3N5 S-scheme heterojunction with oxygen vacancies (OVs) via in-situ solvothermal synthesis (Figure 14a). Benefiting from the efficient separation and transfer of photogenerated charge carriers by S-scheme charge transfer, enriched structural defects, and the close contact interface formed by the in-situ growth, this material remarkably enhanced visible-light photocatalytic degradation efficiencies for TC and Cr(VI) (Figure 14b), whose degradation mechanisms are shown in Figure 14c. The development of S-scheme heterojunctions holds great promise for the design of g-C3N5-based photocatalysts and the applications of g-C3N5-based materials in energy conversion and environmental remediation.

3.2.5. Schottky Junction

A Schottky junction is made from a metal and an n-type semiconductor material, similar to a p-n junction. N-type semiconductors have a higher Fermi energy level than metal, so electrons flow from the semiconductor to the metal. This makes an electron accumulation layer on the metal side and an electron depletion layer on the semiconductor side. This creates an electric field from the semiconductor to the metal, or a “Schottky barrier”, permitting electrons to go from the semiconductor to the metal but not return. The metal plays the role of converging electrons and promotes the effective separation of electrons and holes. In this section, we will talk about three types of Schottky materials for g-C3N5: conventional structures, core-shell structures, and two-dimensional nanosheet structures.
In conventional Schottky connections, metal particles are dispersed or deposited on the semiconductor surface. For example, Zhang et al. [59] fabricated Pt-C3N5 and P-Pt-C3N5 by loading Pt on g-C3N5 through NABH4 reduction and in-situ photo-deposition, respectively. Chen et al. [60] constructed a Ni-C3N5 material by loading Ni onto g-C3N5 using a simple annealing method. The Schottky barrier allows electrons to pass from the g-C3N5 conductor band to Pt (Ni) but not back, thereby preventing the combination of photogenerated electrons and holes in g-C3N5 and enhancing the photocatalytic performance of Pt- and Ni-C3N5.
The core-shell structure is formed by wrapping metal nanoparticles with semiconductor materials, thus forming strong contact between two materials at the interface. Li et al. [61] synthesized HCNs@ATTZ by impregnating 5-amino-1H-tetrazole (5-ATTZ) into hollow carbon nanospheres (HCNs), which were subsequently carbonized at 400 °C to produce HCNs@g-C3N5 (Figure 15a). This is a core-shell-structured Schottky heterojunction in which HCNs are the core and g-C3N5 is the shell (Figure 15b). The difference in Fermi levels produced band bending and an internal electric field. They act similarly to a Schottky heterojunction to promote the separation of electrons and holes.
Two-dimensional (2D) nanosheet materials with chemically active surfaces or edges have a larger specific surface area, which may provide more active sites for photoelectrochemical reactions compared to pristine particles. Meng et al. [62] fabricated the Schottky heterojunction C3N5/RGO containing 2D-structured g-C3N5 and 2D-structured reduced graphene oxide (Figure 15c). Electrons were transported from g-C3N5 to RGO through a Schottky heterojunction, efficiently separating electrons and holes. Moreover, RGO with high electrical conductivity was favorable for electron migration. As a result, the produced C3N5/RGO exhibits good photoreduction for uranium (Figure 15d,e).

4. Environmental and Energy Applications

g-C3N5 has been extensively applied in two major areas: energy conversion (H2 evolution and N2 reduction) and environmental remediation (pollutant degradation, CO2 reduction, NO removal, and peroxide activation). They will be discussed in the subsequent chapters.

4.1. H2 Evolution Reaction (HER)

Solar-driven water-splitting technology has become a research hotspot for sustainable hydrogen manufacturing technology. According to the thermodynamic properties, the process of overall H2O spitting for hydrogen and oxygen evolution is non-spontaneous with a Gibbs free energy of 237.2 kJ·mol−1, which is equal to a potential of 1.23 eV. For the hydrogen evolution reaction (HER), the potential at the semiconductor’s conduction band bottom must be more negative than the hydrogen evolution potential (H+/H2 = 0 eV). For the oxygen evolution reaction (OER), the potential at the valence band top must be more positive than the oxygen evolution potential (H2O/O2 = 1.23 eV). If you want to achieve H2 and O2 evolutions at the same time, the band gap of the semiconductor photocatalyst must be greater than 1.23 eV. For example, Liu et al. [30] synthesized pristine C3N5 nanosheets with a band gap of 1.5 eV, a CB potential of −0.333 eV, and a VB potential of 1.217 eV. The conduction band bottom is more negative than 0 eV, which satisfies the hydrogen evolution conditions. The hydrogen evolution rate of 28.97 μmol·g−1·h−1 was achieved. Although the band gap is greater than 1.23 eV, the valence band potential is less than 1.23 eV (the OER potential), which means OER cannot take place.
In addition, the g-C3N5 functional materials have also been employed for H2 evolution and exhibit excellent performance. For example, NixSy-C3N5 [63], S-Ni(OH)2-C3N5 [64], Pt-C3N5 [59], CD/MoS2/C3N5 [65], CdS/C3N5 [54], CdS/C3N5(CCN) [66], g-C3N5/Zn0.5Cd0.5S [67], NH2-UiO-66/N-CN-2 [52]. The applications of g-C3N5-based photocatalyst in H2 production are summarized in Table 1.

4.2. CO2 Reduction Reaction (CO2RR)

CO2, being the primary byproduct of burning fossil fuels, may considerably contribute to global warming and environmental harm. Photocatalytic CO2 reduction reactions (CO2RR) are a promising technique that converts CO2 to hydrocarbon fuels. It not only lowers CO2 contaminants but also supplies clean fuels such as methane (CH4), methanol (CH3OH), and ethanol (CH3CH2OH) that can replace fossil fuels. However, because the linear CO2 molecule has a relatively stable physical chemistry property with a Gib free energy of 394.28 KJ·mol−1, it is difficult to convert CO2 into C1 compounds (CO, CH4, and CH3OH) and C2 compounds (CH3CH2OH). To break the initial carbon-oxygen bond (750 KJ·mol−1), a tremendous amount of energy is required. Due to the remarkable improvement in visible light utilization, g-C3N5-based materials are expected to be candidates for CO2RR. The photogenerated holes can oxidize H2O into O2 and H+, which H+ can couple with electrons to reduce CO2 to C1 products, followed by coupling C and C to obtain the C2 product. In order to accomplish the above reduction process, the potential at VB must be more positive than that of O2/H2O, and the potential at CB bottom must be more negative than that of C1/CO2.
Adsorption is the key step for CO2 conversion. Cao et al. [29] demonstrated that g-C3N5 (−N = N−) with basicity groups exhibits preferable affinity and notable adsorption selectivity for CO2 by DFT calculation. At 298 K, the adsorption capacities of CO2, CH4, and H2 are 11.8 mmol g−1 at 30 bar, 7.9 mmol g−1 at 50 bar, and 0.75 mmol g−1 at 50 bar, respectively (Figure 16a). It indicated that g-C3N5 is a viable choice for the adsorptive separation of CO2/H2 and CH4/H2, particularly CO2/H2 (Figure 16b). Vinu et al. [19] prepared the ordered mesoporous g-C3N5 (MCN-8E-T) for CO2 adsorption with 5.63 mmol g−1 of capacity at 273 K and 30 bar (Figure 16c). Vinu’s colleagues [27] also investigated MCN-14 materials with different pore sizes, which exhibit 5.6–9.1 mmol g−1 of CO2 adsorption abilities at 0 °C and 30 bar and 14–38% of Faraday efficiencies for CO formation. Except for this conversion from CO2 to CO, Morikawa et al. [69] demonstrated that g-C3N5 (-N = N-) can efficiently catalyze CO2 to CH4 and CH3CH2OH under visible light illumination by DFT calculation, with −0.54 eV (Figure 16d) and −0.61 eV (Figure 16e) of limiting potentials, respectively. Both of them were less than HER free energy (1.18 eV) (Figure 16f), which indicated that g-C3N5 is an effective catalyst for CO2RR due to its strong suppressing effect on HER. Bharati Debnath and coworkers [35] reported that 1% K-doped g-C3N5 (-N = N-) could convert CO2 to CH4 with H+ provided by H2O, displaying 100% selectivity (Figure 17g–i). The keys to improving CO2 absorption and reduction may be a large specific surface area, the suitable size of pores, a more negative conduction potential with strong reduction capacity, and effective charge separation.

4.3. Nitrogen Reduction Reaction (NRR)

Inspired by the process of biological nitrogen fixation, photocatalytic nitrogen reduction for ammonia synthesis has garnered attention. As the g-C3N5 is excited by photons, the generated holes can oxidize H2O into O2 and H+, and electrons may activate N2 with the aid of H+ to generate NH3. Compared with the Haber-Bosch method, this progressive hydrogenation method avoids energy waste and CO2 emissions, which is both energy-saving and environmentally friendly. The potential at the CB bottom needs to be more negative compared to the N2/reduction products (N2/N2·, N2/N2H, N2/N2H2, N2/N2H4, N2/N2H5+, and N2/NH3), while the potential at VB top needs to be more positive compared to that of O2/H2O.
Bharati Debnath [70] designed a g-C3N5/NiCr-LDH heterostructure, a hybrid of g-C3N5 and NiCr-layered double hydroxide. DFT calculations showed its ability to adsorb and activate N2 molecules. The possible adsorption configurations and binding sites, the optimized configurations and adsorption energy (Eads), and the N–N bond lengths of the adsorbed N2 molecules are presented in Figure 17a–d, Figure 17e–h, and Figure 17i–l, respectively. The lesser absorption energy (−0.64 eV/N2) confirmed that the Cr atom is the best adsorption site of the N2 molecule. The increased N–N bond length (1.16 Å) compared to the free N2 molecule (1.078Å) suggests the activation of the N2 molecule. The experimental results also confirmed that the outstanding NH3 yields of g-C3N5/NiCr-LDH were 7.51 and 2.86-folds larger than those of pristine g-C3N5 and NiCr-LDH, respectively (Figure 17m). This is mainly due to the separation of electrons and holes produced by the heterojunction interface between g-C3N5 and NiCr-LDH, which facilitates more electrons for reducing N2 to NH3 (Figure 17n).
B-C3N5, produced by doping boron into g-C3N5, was used to immobilize N2 for NH3 synthesis, as confirmed by in-situ DRIFTS spectra (Figure 17o–p), with an NH3 yield 1.72 times that of pristine g-C3N5 (Figure 17q). B-doping enhances light absorption and electron-hole separation and also provides B-O-H functional groups. The B site is capable of adsorbing and activating nitrogen, while the O-H site supplies the H+ to N2 for NH3 formation [37]. Moreover, Fe, an essential element of nitrogenase, plays an important role in N2 fixation; abundant oxygen vacancies (OVs) generated by W18O49 can adsorb and activate N2, and the narrow band gap of g-C3N5 expands the use of light. Based on these advantages, Li et al. [71] constructed Fe-W18O49/g-C3N5 hybrids for NH3 production at a rate of 131.6 mol·g−1·h−1, which was much superior to that of pristine g-C3N5 or Fe-W18O49.

4.4. NO Removal

Nitrogen oxides (NOx), especially NO, present a grave risk to global ecosystems. Photocatalysis is a viable NO removal technique that can convert NO to NO under visible light illumination. For example, Zhang et al. [60] constructed a Ni-C3N5 photocatalyst for NO removal. Under light irradiation, it increased O2 absorption (Figure 18a) and e/h+ separation, promoting O2 activation and the production of more reactive oxygen species (ROS). In-situ DRIFT spectroscopy was used to investigate the conversion of NO to NO3(Figure 18b). Ion chromatography was used to identify the formation of NO3 (Figure 18c). Degradation experimental results show that 0.1-Ni-C3N5 has the largest NO removal efficiency (about 54%) (Figure 18d). A series of quenching experiments confirmed that the ROS (·O2, 1O2, and ·OH) play an important role in NO removal (Figure 18e). The mechanism of NO removal by g-C3N5-based photocatalyst is further understood, as can be seen in Figure 18f. When photons with sufficient energy irradiate the Ni-C3N5, the electrons leap to the conduction band, forming holes in the valence band. The electrons can activate O2 to generate reactive oxygen species (·O2, 1O2, and ·OH). The h+ and the generated ·O2, 1O2, and ·OH can convert NO to NO3. Moreover, TiO2 (P25)-C3N5 [72] Z-scheme heterojunctions displayed a NO removal rate of 67.1%, which is attributed to the effective separation of photogenerated e and h+ and the reserve of strong redox potentials provided by Z-scheme heterojunctions, promoting more ROS formation. Table 2 presents applications of g-C3N5-based materials for photocatalytic NO removal.

4.5. Pollutant Degradation

Semiconductor materials based on g-C3N5 have been used extensively for the removal of different pollutants from water. Figure 19 summarizes the application of g-C3N5-based materials in the degradation of pollutants, including direct photocatalytic degradation and activated peroxide degradation. We first summarized the direct photocatalytic processes of the g-C3N5-based materials alone. When g-C3N5 is irradiated with photons of sufficient energy, electrons jump to the conduction band and generate holes in the valence band. Electrons oxidize O2 to generate ROS (·O2, 1O2, ·OH), whereas holes oxidize H2O to generate ·OH. The resultant h+, ·O2, 1O2, and ·OH may oxidize organic pollutants to form CO2 and H2O, while electrons can photo-reduce pollutants from high valence to low valence. For example, Cr (VI) and U (VI) can be reduced to Cr (IV) and U (IV), respectively [44,45,62,73]. Meanwhile, the h+ can also oxidize Hg0 to Hg2+ [43].
Moreover, persulfate (PMS/PDS) and sodium percarbonate (SPC) can be activated by g-C3N5-based materials to produce ROS for pollutant degradation. There are both free radicals and non-free radicals in the persulfate activation by g-C3N5-based materials. The free radical process refers to the oxidative degradation of pollutants by active species, such as SO4·, ·OH, and ·O2, produced by activating persulfate; the non-radical process includes the oxidation of pollutants by 1O2, electron transfer between persulfate and pollutants, and the direct oxidation of pollutants by persulfate. For example, Zhu et al. [28] demonstrated that g-C3N5 can activate PMS or PDS with the assistance of visible light to degrade emerging micropollutants. PMS can be activated by g-C3N5 to produce SO4· and ·OH. Meanwhile, electrons in the g-C3N5 conduction band can react with O2 to produce ·O2 and 1O2. The above species (h+, ·O2, SO4·, 1O2, and ·OH) are jointly involved in the degradation process of EMs. Ding et al. [39] demonstrated PDA-g-CN-1.0 obtained by coating g-C3N5 with polydopamine could effectively activate PMS without light to generate high-potential C-PMS* complexes, which could completely oxidize SMX within 5 min. Fang et al. [32] demonstrated that Co-C3N5 can activate PMS without light to completely degrade PCB28 within 30 min, which was significantly superior to the traditional metal-based activation. It was found that sulfate radical was the main active species for PCB28 degradation, which was generated from the decomposition of PMS activated by Co-N4. In contrast to this conclusion, Zhu et al. [33] discovered that high-valent cobalt oxides play an important role in SMX degradation by another Co-C3N5 activating PMS with the help of visible light. This difference may be caused by the combination of photocatalysis and the transition metal cobalt under visible light irradiation.
Sodium percarbonate (SPC, Na2CO3·1.5H2O2), as a solid carrier of H2O2, not only has the properties of H2O2, but also has the possibility of producing carbonate radicals. Due to its easy transportation, storage, and handling properties, it has been used as a substitute for H2O2 in AOPs. Zhang et al. [67] employed ultrathin-C3N5 and SPC coupling photocatalytic techniques to degrade sulfamethoxazole, arriving at a removal rate of 93.97% in 120 min. The main processes are as follows: (1) the separation of e and h+; (2) the formation of ROS (·OH, ·O2, and 1O2); and (3) the degradation of pollutants by ROS (·OH, ·O2, 1O2, ·CO3, and h+). This work expands the application of g-C3N5-based materials in AOPs. Table 3 summarizes the recent developments in the photo-degradation and photo-reduction of pollutants by g-C3N5-based photocatalyst.

5. Conclusions and Future Outlook

In summary, this review presents an overview of the synthesis, functionalization, and applications of g-C3N5-based photocatalysts. Tremendous advancements have been witnessed during the past few years. Nevertheless, there are still some challenges and opportunities to be tackled.
Firstly, most reported g-C3N5-based photocatalysts are limited to the laboratory scale. To reach an industrial scale, time-saving and environmentally friendly processes for the synthesis of g-C3N5-based materials must be investigated. The new technology for the synthesis of g-C3N5 should be rapid, cost-effective, and time-efficient. Toxic templates, eco-damaging etchants, and complicated procedures should be avoided. Additionally, material recycling and sustainability are crucial elements for implementing industrial applications at scale. Combining g-C3N5 with narrower bandgap semiconductors can increase the efficiency of exploiting the full spectrum of sunlight; constructing suited heterojunctions can facilitate the separation of photoinduced carriers while preserving a high redox capacity; and coupling with magnetic materials can promote quick recycling.
Photocatalytic reactions heavily depend on separated electrons and holes. In order to comprehend the photocatalytic mechanism of the detection and characterization of electrons, holes, and reaction intermediates are vital. However, it is difficult to identify the transiently changing electrons, holes, and reaction intermediates using conventional characterization techniques. Therefore, more progressive technologies, such as in situ XPS, in situ IR spectra, and time-resolved terahertz spectra, may be employed to unveil the transitory processes of reactions.
As model pollutants for assessing photocatalytic activity, MB, MO, and RhB are often used. However, their photosensitivity and adsorption affect degradation efficiency, which is not an independent contribution of the photocatalysts. To accurately determine the photocatalytic efficacy of g-C3N5-based materials, it is necessary to employ other model pollutants, including medicines, pesticides, and phenols. In addition to elucidating the efficiency and processes of degradation, the biotoxicity and other possible dangers of degradation byproducts should be thoroughly explored.
With the further advancement of experimental research and theoretical calculations, the underlying photocatalysis process will be better understood, and the solar-to-energy conversion bottleneck might be overcome. Massive commercial applications of g-C3N5-based photocatalysts will foster renewable revolutions in the fields of energy and environmental engineering.

Author Contributions

Conceptualization, C.Z. and J.Z.; methodology, C.Z. and J.Z.; writing—original draft preparation, J.L.; writing—review and editing, J.L. and S.W.; visualization, J.L. and S.W.; supervision, J.L. and S.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hasija, V.; Nguyen, V.-H.; Kumar, A.; Raizada, P.; Krishnan, V.; Khan, A.A.P.; Singh, P.; Lichtfouse, E.; Wang, C.; Thi Huong, P. Advanced activation of persulfate by polymeric g-C3N4 based photocatalysts for environmental remediation: A review. J. Hazard. Mater. 2021, 413, 125324. [Google Scholar] [CrossRef] [PubMed]
  2. Behera, P.; Subudhi, S.; Tripathy, S.P.; Parida, K. MOF derived nano-materials: A recent progress in strategic fabrication, characterization and mechanistic insight towards divergent photocatalytic applications. Coord. Chem. Rev. 2022, 456, 214392. [Google Scholar] [CrossRef]
  3. Wang, Y.; Ding, Z.; Arif, N.; Jiang, W.-C.; Zeng, Y.-J. 2D material based heterostructures for solar light driven photocatalytic H2 production. Mater. Adv. 2022, 3, 3389–3417. [Google Scholar] [CrossRef]
  4. Lakhi, K.S.; Park, D.H.; Al-Bahily, K.; Cha, W.; Viswanathan, B.; Choy, J.H.; Vinu, A. Mesoporous carbon nitrides: Synthesis, functionalization, and applications. Chem. Soc. Rev. 2017, 46, 72–101. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, X.; Ye, Y.; Sun, J.; Li, Z.; Ping, J.; Sun, X. Recent Advances in g-C3N4-Based Photocatalysts for Pollutant Degradation and Bacterial Disinfection: Design Strategies, Mechanisms, and Applications. Small 2022, 18, 202105089. [Google Scholar]
  6. Acharya, R.; Pati, S.; Parida, K. A review on visible light driven spinel ferrite-g-C3N4 photocatalytic systems with enhanced solar light utilization. J. Mol. Liq. 2022, 357, 119105. [Google Scholar] [CrossRef]
  7. Wang, S.J.; Zhang, J.Q.; Li, B.; Sun, H.Q.; Wang, S.B.; Duan, X.G. Morphology-dependent photocatalysis of graphitic carbon nitride for sustainable remediation of aqueous pollutants: A mini review. J. Environ. Chem. Eng. 2022, 10, 107438. [Google Scholar] [CrossRef]
  8. Safian, M.T.U.; Umar, K.; Ibrahim, M.N.M. Synthesis and scalability of graphene and its derivatives: A journey towards sustainable and commercial material. J. Clean Prod. 2021, 318, 128603. [Google Scholar] [CrossRef]
  9. Lopez-Salas, N.; Albero, J. CxNy: New Carbon Nitride Organic Photocatalysts. Front. Mater. 2021, 8, 772200. [Google Scholar] [CrossRef]
  10. Chen, R.; Ren, Z.; Liang, Y.; Zhang, G.; Dittrich, T.; Liu, R.; Liu, Y.; Zhao, Y.; Pang, S.; An, H.; et al. Spatiotemporal imaging of charge transfer in photocatalyst particles. Nature 2022, 610, 296–301. [Google Scholar] [CrossRef]
  11. Wang, Q.; Fang, Z.; Zhang, W.; Zhang, D. High-Efficiency g-C3N4 Based Photocatalysts for CO2 Reduction: Modification Methods. Adv. Fiber Mater. 2022, 4, 342–360. [Google Scholar] [CrossRef]
  12. Wang, S.; Zhang, J.; Li, B.; Sun, H.; Wang, S. Engineered Graphitic Carbon Nitride-Based Photocatalysts for Visible-Light-Driven Water Splitting: A Review. Energy Fuels 2021, 35, 6504–6526. [Google Scholar] [CrossRef]
  13. Balakrishnan, A.; Chinthala, M. Comprehensive review on advanced reusability of g-C3N4 based photocatalysts for the removal of organic pollutants. Chemosphere 2022, 297, 134190. [Google Scholar] [CrossRef]
  14. He, F.; Liu, X.; Zhao, X.; Zhang, J.; Dong, P.; Zhang, Y.; Zhao, C.; Sun, H.; Duan, X.; Wang, S.; et al. Manipulation of n → π* electronic transitions via implanting thiophene rings into two-dimensional carbon nitride nanosheets for efficient photocatalytic water purification. J. Mater. Chem. A 2022, 10, 20559–20570. [Google Scholar] [CrossRef]
  15. Pérez-Álvarez, D.T.; Brown, J.; Elgohary, E.A.; Mohamed, Y.M.A.; El Nazer, H.A.; Davies, P.; Stafford, J. Challenges surrounding nanosheets and their application to solar-driven photocatalytic water treatment. Mater. Adv. 2022, 3, 4103–4131. [Google Scholar] [CrossRef]
  16. Rajendran, S.; Chellapandi, T.; UshaVipinachandran, V.; Venkata Ramanaiah, D.; Dalal, C.; Sonkar, S.K.; Madhumitha, G.; Bhunia, S.K. Sustainable 2D Bi2WO6/g-C3N5 heterostructure as visible light-triggered abatement of colorless endocrine disruptors in wastewater. Appl. Surf. Sci. 2022, 577, 151809. [Google Scholar] [CrossRef]
  17. Kumar, P.; Vahidzadeh, E.; Thakur, W.K.; Kar, P.; Alam, K.M.; Goswami, A.; Mandi, N.; Cui, K.; Bernard, G.M.; Michaelis, V.K.; et al. C3N5: A Low Bandgap Semiconductor Containing an Azo-Linked Carbon Nitride Framework for Photocatalytic, Photovoltaic and Adsorbent Applications. J. Am. Chem. Soc. 2019, 141, 5415–5436. [Google Scholar] [CrossRef] [PubMed]
  18. Mane, G.P.; Talapaneni, S.N.; Lakhi, K.S.; Ilbeygi, H.; Ravon, U.; Al-Bahily, K.; Mori, T.; Park, D.H.; Vinu, A. Highly Ordered Nitrogen-Rich Mesoporous Carbon Nitrides and Their Superior Performance for Sensing and Photocatalytic Hydrogen Generation. Angew.Chem. Int. Ed. 2017, 56, 8481–8485. [Google Scholar] [CrossRef] [Green Version]
  19. Park, D.H.; Lakhi, K.S.; Ramadass, K.; Kim, M.K.; Talapaneni, S.N.; Joseph, S.; Ravon, U.; Al-Bahily, K.; Vinu, A. Energy Efficient Synthesis of Ordered Mesoporous Carbon Nitrides with a High Nitrogen Content and Enhanced CO2 Capture Capacity. Chem.Eur. J. 2017, 23, 10753–10757. [Google Scholar] [CrossRef] [PubMed]
  20. Wang, H.; Li, M.; Lu, Q.; Cen, Y.; Zhang, Y.; Yao, S. A Mesoporous Rod-like g-C3N5 Synthesized by Salt-Guided Strategy: As a Superior Photocatalyst for Degradation of Organic Pollutant. ACS Sustain. Chem. Eng. 2019, 7, 625–631. [Google Scholar] [CrossRef]
  21. Kim, I.Y.; Kim, S.; Jin, X.; Premkumar, S.; Chandra, G.; Lee, N.-S.; Mane, G.P.; Hwang, S.-J.; Umapathy, S.; Vinu, A. Ordered Mesoporous C3N5 with a Combined Triazole and Triazine Framework and Its Graphene Hybrids for the Oxygen Reduction Reaction (ORR). Angew. Chem. Int.Ed. 2018, 57, 17135–17140. [Google Scholar] [CrossRef]
  22. Liu, X.; Xing, S.; Xu, Y.; Chen, R.; Lin, C.; Guo, L. 3-Amino-1,2,4-triazole-derived graphitic carbon nitride for photodynamic therapy. Spectrochim. Acta. A. Mol. Biomol. Spectrosc. 2021, 250, 119363. [Google Scholar] [CrossRef]
  23. Yang, H.; Zhou, Q.; Fang, Z.; Li, W.; Zheng, Y.; Ma, J.; Wang, Z.; Zhao, L.; Liu, S.; Shen, Y.; et al. Carbon nitride of five-membered rings with low optical bandgap for photoelectrochemical biosensing. Chem 2021, 7, 2708–2721. [Google Scholar] [CrossRef]
  24. Zeng, X.; Hou, M.; Zhu, P.; Yuan, M.; Ouyang, S.; Lu, Q.; Zhao, C.; Wang, H.; Du, F.; Zeng, G.; et al. g-C3N5-dots as fluorescence probes prepared by an alkali-assisted hydrothermal method for cell imaging. RSC Adv. 2022, 12, 26476–26484. [Google Scholar] [CrossRef]
  25. Wang, H.Y.; Li, M.X.; Li, H.; Lu, Q.J.; Zhang, Y.Y.; Yao, S.Z. Porous graphitic carbon nitride with controllable nitrogen vacancies: As promising catalyst for enhanced degradation of pollutant under visible light. Mater. Des. 2019, 162, 210–218. [Google Scholar] [CrossRef]
  26. Ruban, S.M.; Sathish, C.I.; Ramadass, K.; Joseph, S.; Kim, S.; Dasireddy, V.D.B.C.; Kim, I.Y.; Al-Muhtaseb, A.H.; Sugi, Y.; Vinu, A. Ordered Mesoporous Carbon Nitrides with Tuneable Nitrogen Contents and Basicity for Knoevenagel Condensation. Chemcatchem 2021, 13, 468–474. [Google Scholar] [CrossRef]
  27. Kim, S.; Singh, G.; Sathish, C.I.; Panigrahi, P.; Daiyan, R.; Lu, X.; Sugi, Y.; Kim, I.Y.; Vinu, A. Tailoring the Pore Size, Basicity, and Binding Energy of Mesoporous C3N5 for CO2 Capture and Conversion. Chem.-Asian J. 2021, 16, 3999–4005. [Google Scholar] [CrossRef]
  28. Zhang, J.L.; Jing, B.H.; Tang, Z.Y.; Ao, Z.M.; Xia, D.H.; Zhu, M.S.; Wang, S.B. Experimental and DFT insights into the visible-light driving metal-free C3N5 activated persulfate system for efficient water purification. Appl. Catal. B-Environ. 2021, 289, 120023. [Google Scholar] [CrossRef]
  29. Huang, L.; Liu, Z.; Chen, W.; Cao, D.; Zheng, A. Two-dimensional graphitic C3N5 materials: Promising metal-free catalysts and CO2 adsorbents. J. Mater. Chem. A 2018, 6, 7168–7174. [Google Scholar] [CrossRef]
  30. Liu, T.; Yang, G.; Wang, W.; Wang, C.; Wang, M.; Sun, X.; Xu, P.; Zhang, J. Preparation of C3N5 nanosheets with enhanced performance in photocatalytic methylene blue (MB) degradation and H2-evolution from water splitting. Environ. Res. 2020, 188, 109741. [Google Scholar] [CrossRef]
  31. Quan, C.Y.; Xiao, S.S.; Yi, Y.W.; Sun, D.Z.; Ji, S.L.; Zhou, S.; Yang, J.P.; Niu, X.H.; Li, X.A. Explore the underlying mechanism of graphitic C3N5-hosted single-atom catalyst for electrocatalytic nitrogen fixation. Int. J. Hydrogen Energy 2022, 47, 22035–22044. [Google Scholar] [CrossRef]
  32. Li, T.; Cui, P.X.; Wang, X.L.; Liu, C.; Zeng, Y.; Fang, G.D.; Zhao, Y.; Gao, J.; Wang, Y.J.; Zhou, D.M. Efficient activation of peroxymonosulfate by C3N5 doped with cobalt for organic contaminant degradation. Environ. Sci.-Nano 2022, 9, 2534–2547. [Google Scholar] [CrossRef]
  33. He, P.; Gu, C.; Tang, B.; Zhou, Y.; Gan, M.; Zhu, J. Expeditious degradation of SMX by high-valent cobalt-oxo species derived from cobalt-doped C3N5-activated peroxymonosulfate with the assistance of visible light. Sep. Purif. Technol. 2022, 301, 122009. [Google Scholar] [CrossRef]
  34. Gujral, H.S.; Singh, G.; Yang, J.H.; Sathish, C.I.; Yi, J.; Karakoti, A.; Fawaz, M.; Ramadass, K.; Al-Muhtaseb, A.a.H.; Yu, X.; et al. Mesoporous titanium carbonitride derived from mesoporous C3N5 for highly efficient hydrogen evolution reaction. Carbon 2022, 195, 9–18. [Google Scholar] [CrossRef]
  35. Debnath, B.; Singh, S.; Hossain, S.M.; Krishnamurthy, S.; Polshettiwar, V.; Ogale, S. Visible Light-Driven Highly Selective CO2 Reduction to CH4 Using Potassium-Doped g-C3N5. Langmuir 2022, 38, 3139–3148. [Google Scholar] [CrossRef]
  36. Hu, C.; Lin, Y.-H.; Yoshida, M.; Ashimura, S. Influence of Phosphorus Doping on Triazole-Based g-C3N5 Nanosheets for Enhanced Photoelectrochemical and Photocatalytic Performance. ACS Appl. Mater. Interfaces 2021, 13, 24907–24915. [Google Scholar] [CrossRef]
  37. Li, K.; Cai, W.; Zhang, Z.; Xie, H.; Zhong, Q.; Qu, H. Boron doped C3N5 for photocatalytic nitrogen fixation to ammonia: The key role of boron in nitrogen activation and mechanism. Chem. Eng. J. 2022, 435, 135017. [Google Scholar] [CrossRef]
  38. Sun, D.Z.; Zhang, X.M.; Shi, A.Q.; Quan, C.Y.; Xiao, S.S.; Ji, S.L.; Zhou, Z.B.; Li, X.A.; Chi, F.F.; Niu, X.H. Metal-free boron doped g-C3N5 catalyst: Efficient doping regulatory strategy for photocatalytic water splitting. Appl. Surf. Sci. 2022, 601, 154186. [Google Scholar] [CrossRef]
  39. Cai, S.; Zuo, X.; Zhao, H.; Yang, S.; Chen, R.; Chen, L.; Zhang, R.; Ding, D.; Cai, T. Evaluation of N-doped carbon for the peroxymonosulfate activation and removal of organic contaminants from livestock wastewater and groundwater. J. Mater. Chem. A 2022, 10, 9171–9183. [Google Scholar] [CrossRef]
  40. Che, H.; Wang, J.; Gao, X.; Chen, J.; Wang, P.; Liu, B.; Ao, Y. Regulating directional transfer of electrons on polymeric g-C3N5 for highly efficient photocatalytic H2O2 production. J. Colloid Interface Sci. 2022, 627, 739–748. [Google Scholar] [CrossRef]
  41. Li, M.; Lu, Q.; Liu, M.; Yin, P.; Wu, C.; Li, H.; Zhang, Y.; Yao, S. Photoinduced Charge Separation via the Double-Electron Transfer Mechanism in Nitrogen Vacancies g-C3N5/BiOBr for the Photoelectrochemical Nitrogen Reduction. ACS Appl. Mater. Interfaces 2020, 12, 38266–38274. [Google Scholar] [CrossRef]
  42. Cai, Z.; Huang, Y.; Ji, H.; Liu, W.; Fu, J.; Sun, X. Type-II surface heterojunction of bismuth-rich Bi4O5Br2 on nitrogen-rich g-C3N5 nanosheets for efficient photocatalytic degradation of antibiotics. Sep. Purif. Technol. 2022, 280, 119772. [Google Scholar] [CrossRef]
  43. Chu, W.; Li, S.; Zhou, H.; Shi, M.; Zhu, J.; He, P.; Liu, W.; Wu, M.; Wu, J.; Yan, X. A novel bionic flower-like Z-scheme Bi4O5I2/g-C3N5 heterojunction with I/I3 active sites and π-conjugated structure for increasing photocatalytic oxidation of elemental mercury. J. Environ. Chem. Eng. 2022, 10, 108623. [Google Scholar] [CrossRef]
  44. Li, S.J.; Cai, M.J.; Liu, Y.P.; Zhang, J.L.; Wang, C.C.; Zang, S.H.; Li, Y.J.; Zhang, P.; Li, X. In situ construction of a C3N5 nanosheet/Bi2WO6 nanodot S-scheme heterojunction with enhanced structural defects for the efficient photocatalytic removal of tetracycline and Cr(vi). Inorg. Chem. Front. 2022, 9, 2479–2497. [Google Scholar] [CrossRef]
  45. Li, S.; Wang, C.; Cai, M.; Liu, Y.; Dong, K.; Zhang, J. Designing oxygen vacancy mediated bismuth molybdate (Bi2MoO6)/N-rich carbon nitride (C3N5) S-scheme heterojunctions for boosted photocatalytic removal of tetracycline antibiotic and Cr(VI): Intermediate toxicity and mechanism insight. J. Colloid Interface Sci. 2022, 624, 219–232. [Google Scholar] [CrossRef]
  46. Vadivel, S.; Hariganesh, S.; Paul, B.; Mamba, G.; Puviarasu, P. Highly active novel CeTi2O6/g-C3N5 photocatalyst with extended spectral response towards removal of endocrine disruptor 2, 4-dichlorophenol in aqueous medium. Colloid Surf. A-Physicochem. Eng. Asp. 2020, 592, 124583. [Google Scholar] [CrossRef]
  47. Yin, H.; Cao, Y.; Fan, T.; Zhang, M.; Yao, J.; Li, P.; Chen, S.; Liu, X. In situ synthesis of Ag3PO4/C3N5 Z-scheme heterojunctions with enhanced visible-light-responsive photocatalytic performance for antibiotics removal. Sci. Total Environ. 2021, 754, 141926. [Google Scholar] [CrossRef]
  48. Li, G.; Zeng, G.; Chen, Z.; Hong, J.; Ji, X.; Lan, Z.; Tan, X.; Li, M.; Hu, X.; Tang, C. In Situ Coupling Carbon Defective C3N5 Nanosheet with Ag2CO3 for Effective Degradation of Methylene Blue and Tetracycline Hydrochloride. Nanomaterials 2022, 12, 2701. [Google Scholar] [CrossRef]
  49. Vadivel, S.; Hariganesh, S.; Paul, B.; Rajendran, S.; Habibi-Yangjeh, A.; Maruthamani, D.; Kumaravel, M. Synthesis of novel AgCl loaded g-C3N5 with ultrahigh activity as visible light photocatalyst for pollutants degradation. Chem. Phys. Lett. 2020, 738, 136862. [Google Scholar] [CrossRef]
  50. Vadivel, S.; Fujii, M.; Rajendran, S. Facile synthesis of broom stick like FeOCl/g-C3N5 nanocomposite as novel Z-scheme photocatalysts for rapid degradation of pollutants. Chemosphere 2022, 307, 135716. [Google Scholar] [CrossRef]
  51. Wang, R.; Zhang, K.; Zhong, X.; Jiang, F. Z-scheme LaCoO3/C3N5 for efficient full-spectrum light-simulated solar photocatalytic hydrogen generation. RSC Adv. 2022, 12, 24026–24036. [Google Scholar] [CrossRef] [PubMed]
  52. Wu, B.; Sun, T.; Liu, N.; Lu, L.; Zhang, R.; Shi, W.; Cheng, P. Modulation of Z-Scheme Heterojunction Interface between Ultrathin C3N5 Nanosheets and Metal-Organic Framework for Boosting Photocatalysis. ACS Appl. Mater. Interfaces 2022, 14, 26742–26751. [Google Scholar] [CrossRef] [PubMed]
  53. Alam, K.M.; Jensen, C.E.; Kumar, P.; Hooper, R.W.; Bernard, G.M.; Patidar, A.; Manuel, A.P.; Amer, N.; Palmgren, A.; Purschke, D.N.; et al. Photocatalytic Mechanism Control and Study of Carrier Dynamics in CdS@C3N5 Core-Shell Nanowires. ACS Appl. Mater. Interfaces 2021, 13, 47418–47439. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, B.; Qiao, H.; Guan, P.; Yang, B.; Liu, B. Fabrication of CdS/C3N5 photocatalyst for enhanced H2 production. Compos. Interfaces 2022, 2076363. [Google Scholar] [CrossRef]
  55. Yi, F.; Liu, J.; Liang, G.; Xiao, X.; Wang, H. Insight into the enhanced degradation mechanism of g-C3N4/g-C3N5 heterostructures through photocatalytic molecular oxygen activation in Van der Waals junction and excitation. J. Alloys Compd. 2022, 905, 164064. [Google Scholar] [CrossRef]
  56. Meng, Q.; Yang, X.; Wu, L.; Chen, T.; Li, Y.; He, R.; Zhu, W.; Zhu, L.; Duan, T. Metal-free 2D/2D C3N5/GO nanosheets with customized energy-level structure for radioactive nuclear wastewater treatment. J. Hazard. Mater. 2022, 422, 126912. [Google Scholar] [CrossRef]
  57. Wang, L.X.; Bie, C.B.; Yu, J.G. Challenges of Z-scheme photocatalytic mechanisms. Trends Chem. 2022, 4, 973–983. [Google Scholar] [CrossRef]
  58. Zhang, L.; Zhang, J.; Yu, H.; Yu, J. Emerging S-Scheme Photocatalyst. Adv. Mater. 2022, 34, 2107668. [Google Scholar] [CrossRef]
  59. Peng, C.; Han, L.X.; Huang, J.M.; Wang, S.Y.; Zhang, X.H.; Chen, H. Comprehensive investigation on robust photocatalytic hydrogen production over C3N5. Chin. J. Catal. 2022, 43, 410–420. [Google Scholar] [CrossRef]
  60. Han, L.X.; Huang, J.M.; Zhan, J.; Zhang, X.H.; Wang, S.Y.; Chen, H. Construction of Highly Dispersed Ni Sites on N-rich Carbon Nitride for Enhanced Photocatalytic NO Removal. Adv. Sustain. Syst. 2022, 2200009. [Google Scholar] [CrossRef]
  61. Li, Q.; Song, S.; Mo, Z.; Zhang, L.; Qian, Y.; Ge, C. Hollow carbon nanospheres@graphitic C3N5 heterostructures for enhanced oxygen electroreduction. Appl. Surf. Sci. 2022, 579, 152006. [Google Scholar] [CrossRef]
  62. Wu, L.; Yang, X.; Chen, T.; Li, Y.; Meng, Q.; Zhu, L.; Zhu, W.; He, R.; Duan, T. Three-dimensional C3N5 /RGO aerogels with enhanced visible-light response and electron-hole separation efficiency for photocatalytic uranium reduction. Chem. Eng. J. 2022, 427, 131773. [Google Scholar] [CrossRef]
  63. Han, L.X.; Peng, C.; Huang, J.M.; Sun, L.H.; Wang, S.Y.; Zhang, X.H.; Yang, Y. Noble-Metal-Free NixSy-C3N5 Hybrid Nanosheet with Highly Efficient Photocatalytic Performance. Catalysts 2021, 11, 1089. [Google Scholar] [CrossRef]
  64. Han, L.; Peng, C.; Huang, J.; Wang, S.; Zhang, X.; Chen, H.; Yang, Y. Ultrafast synthesis of near-zero-cost S-doped Ni(OH)2 on C3N5 under ambient conditions with enhanced photocatalytic activity. RSC Adv. 2021, 11, 36166–36173. [Google Scholar] [CrossRef]
  65. Teng, M.; Shi, J.; Qi, H.; Shi, C.; Wang, W.; Kang, F.; Eqi, M.; Huang, Z. Effective enhancement of electron migration and photocatalytic performance of nitrogen-rich carbon nitride by constructing fungal carbon dot/molybdenum disulfide cocatalytic system. J. Colloid Interface Sci. 2022, 609, 592–605. [Google Scholar] [CrossRef]
  66. Shi, J.; Wang, W.; Teng, M.; Kang, F.; E′qi, M.; Huang, Z. Biomimetic nitrogen-rich photocatalyst based on cadmium sulfide for photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2022, 608, 954–962. [Google Scholar] [CrossRef]
  67. Jin, Z.; Wei, T.; Huang, J.; Li, F.; Yang, Y.; Xu, L. Fabrication of direct Z-scheme heterojunction between Zn0.5Cd0.5S and N-rich graphite carbon nitride for boosted H2 production. Int. J. Hydrogen Energy 2020, 45, 22711–22721. [Google Scholar] [CrossRef]
  68. Liu, D.; Yao, J.; Chen, S.T.; Zhang, J.; Li, R.J.; Peng, T.Y. Construction of rGO-coupled C3N4/C3N5 2D/2D Z-scheme heterojunction to accelerate charge separation for efficient visible light H2 evolution. Appl. Catal. B-Environ. 2022, 318, 121822. [Google Scholar] [CrossRef]
  69. Wang, Y.; Thanh Ngoc, P.; Tian, Y.; Morikawa, Y.; Yan, L. Density functional theory study on a nitrogen-rich carbon nitride material C3N5 as photocatalyst for CO2 reduction to C1 and C2 products. J. Colloid Interface Sci. 2021, 585, 740–749. [Google Scholar] [CrossRef]
  70. Debnath, B.; Hossain, S.M.; Sadhu, A.; Singh, S.; Polshettiwar, V.; Ogale, S. Construction of a 2D/2D g-C3N5/NiCr-LDH Heterostructure to Boost the Green Ammonia Production Rate under Visible Light Illumination. ACS Appl. Mater. Interfaces 2022, 14, 37076–37087. [Google Scholar] [CrossRef]
  71. Wang, L.; Li, M.; Zhang, Q.; Li, F.; Xu, L. Constructing electron transfer pathways and active centers over W18O49 nanowires by doping Fe3+ and incorporating g-C3N5 for enhanced photocatalytic nitrogen fixation. Inorg. Chem. Front. 2021, 8, 3566–3575. [Google Scholar] [CrossRef]
  72. Zhang, J.L.; Tao, H.C.; Wu, S.S.; Yang, J.L.; Zhu, M.S. Enhanced durability of nitric oxide removal on TiO2 (P25) under visible light: Enabled by the direct Z-scheme mechanism and enhanced structure defects through coupling with C3N5. Appl. Catal. B-Environ. 2021, 296, 120372. [Google Scholar] [CrossRef]
  73. Wang, Z.; Li, W.; Wu, L.; Wang, Z.; Cao, Y.; Cheng, J.; Chen, G.; Zhao, Q.; Jiang, M.; Chen, Z.; et al. Nitrogen-rich carbon nitrogen polymers for enhancing the sorption of uranyl. Chin. Chem. Lett. 2022, 33, 3468–3473. [Google Scholar] [CrossRef]
  74. Zhu, D.; Huang, Z.; Wang, H.; Lu, Q.; Ruan, G.; Zhao, C.; Du, F. Sustainable and reusable electrospun g-C3N5/MIL-101(Fe)/poly(acrylonitrile-co-maleic acid) nanofibers for photocatalytic degradation of emerging pharmaceutical pollutants. New J. Chem. 2022, 46, 11840–11850. [Google Scholar] [CrossRef]
  75. Wei, W.; Gong, H.; Sheng, L.; Wu, H.; Zhu, S.; Feng, L.; Li, X.; You, W. Highly efficient photocatalytic activity and mechanism of novel Er3+ and Tb3+ co-doped BiOBr/g-C3N5 towards sulfamethoxazole degradation. Ceram. Int. 2021, 47, 24062–24072. [Google Scholar] [CrossRef]
  76. He, X.; Zhu, D. In situ solvothermal method of C3N5@NH2-MIL-125 composites with enhanced visible-light photocatalytic performance. J. Mater. Sci. Mater. Electron. 2021, 33, 388–398. [Google Scholar] [CrossRef]
  77. Saravanakumar, K.; Maheskumar, V.; Yea, Y.; Yoon, Y.; Muthuraj, V.; Park, C.M. 2D/2D nitrogen-rich graphitic carbon nitride coupled Bi2WO6 S-scheme heterojunction for boosting photodegradation of tetracycline: Influencing factors, intermediates, and insights into the mechanism. Compos. Pt. B-Eng. 2022, 234, 109726. [Google Scholar] [CrossRef]
  78. Vadivel, S.; Fujii, M.; Rajendran, S. Novel S-scheme 2D/2D Bi4O5Br2 nanoplatelets/g-C3N5 heterojunctions with enhanced photocatalytic activity towards organic pollutants removal. Environ. Res. 2022, 213, 113736. [Google Scholar] [CrossRef]
  79. Ma, C.; Yu, Z.; Wei, J.; Tan, C.; Yang, X.; Wang, T.; Yu, G.; Zhang, C.; Li, X. Metal-free ultrathin C3N5 photocatalyst coupling sodium percarbonate for efficient sulfamethoxazole degradation. Appl. Catal. B-Environ. 2022, 319, 121951. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the g-C3N5 synthesis process by the thermal polymerization of different precursors, such as 5-amino-1H-tetrazole, 3-amino-1,2,4-triazole, and melamine. (a) g-C3N5 with triazine and triazole; (b) g-C3N5 with terminal triazole; (c) g-C3N5 with heptazine linked by azo.
Figure 1. Schematic illustration of the g-C3N5 synthesis process by the thermal polymerization of different precursors, such as 5-amino-1H-tetrazole, 3-amino-1,2,4-triazole, and melamine. (a) g-C3N5 with triazine and triazole; (b) g-C3N5 with terminal triazole; (c) g-C3N5 with heptazine linked by azo.
Nanomaterials 13 00499 g001
Figure 2. Theoretical and experimental results of g-C3N5 containing one-triazole and two-triazine: (a) Optimized geometric structure (gray for C atoms and blue for N atoms); (b) 2D charge density distribution; (c) Phonon DOS; (d) 3D charge density (yellow regions: electron-rich areas, cyan regions: electron-deficient areas); (e) Carbon K-edge (left) and nitrogen K-edge (right) NEXAFS spectra (the black line indicates MCN-11); (f) FT-IR spectra (the black line indicates MCN-11); (g) g-C3N5 containing 1-NH2/NH-1,2,4-triazole units; (h) g-C3N4 containing 2-NH2/NH-1,3,5-triazine units. (af) Adapted with permission from [21]. Copyright Wiley-VCH Verlag GmbH and Co. KGaA, Weinheim, 2018. (gh) Adapted with permission from [26]. Copyright WileyVCH GmbH, 2020.
Figure 2. Theoretical and experimental results of g-C3N5 containing one-triazole and two-triazine: (a) Optimized geometric structure (gray for C atoms and blue for N atoms); (b) 2D charge density distribution; (c) Phonon DOS; (d) 3D charge density (yellow regions: electron-rich areas, cyan regions: electron-deficient areas); (e) Carbon K-edge (left) and nitrogen K-edge (right) NEXAFS spectra (the black line indicates MCN-11); (f) FT-IR spectra (the black line indicates MCN-11); (g) g-C3N5 containing 1-NH2/NH-1,2,4-triazole units; (h) g-C3N4 containing 2-NH2/NH-1,3,5-triazine units. (af) Adapted with permission from [21]. Copyright Wiley-VCH Verlag GmbH and Co. KGaA, Weinheim, 2018. (gh) Adapted with permission from [26]. Copyright WileyVCH GmbH, 2020.
Nanomaterials 13 00499 g002
Figure 3. The characterization of g-C3N5: (a) XRD pattern; (b) XPS survey; (c) high-resolution C1s spectra; (d) high-resolution N1s XPS spectra for g-C3N5; (e) UV–vis DRS; (f) Tauc’s plot; (g) PL spectra; (h) ESR spectra for g-C3N5 and g-C3N4; (i) and (j) The orbitals for g-C3N4 and g-C3N5. Adapted with permission from [28]. Copyright Elsevier B.V., 2021.
Figure 3. The characterization of g-C3N5: (a) XRD pattern; (b) XPS survey; (c) high-resolution C1s spectra; (d) high-resolution N1s XPS spectra for g-C3N5; (e) UV–vis DRS; (f) Tauc’s plot; (g) PL spectra; (h) ESR spectra for g-C3N5 and g-C3N4; (i) and (j) The orbitals for g-C3N4 and g-C3N5. Adapted with permission from [28]. Copyright Elsevier B.V., 2021.
Nanomaterials 13 00499 g003
Figure 4. Structural characterization and applications of MCN-8: (a) Low-angle powder XRD pattern (inset: wide-angle powder XRD pattern); (b) N2 adsorption-desorption isotherm; (c) Carbon K-edge and nitrogen K–edge NEXAFS spectra (red line for MCN-8, blue line for g-C3N4, and the inset for the proposed molecular structure); (d) TEM image; (e) FE-SEM images; (f) High-resolution XPS spectra: C1s and N1s; (g) EEL spectrum; (h) UV–Vis absorption spectrum; (i) H2 evolution amount in visible light; (j) Gas adsorption detected by the QCM sensor. Adapted with permission from [18]. Copyright Wiley-VCH Verlag GmbH and Co. KGaA, Weinheim, 2017.
Figure 4. Structural characterization and applications of MCN-8: (a) Low-angle powder XRD pattern (inset: wide-angle powder XRD pattern); (b) N2 adsorption-desorption isotherm; (c) Carbon K-edge and nitrogen K–edge NEXAFS spectra (red line for MCN-8, blue line for g-C3N4, and the inset for the proposed molecular structure); (d) TEM image; (e) FE-SEM images; (f) High-resolution XPS spectra: C1s and N1s; (g) EEL spectrum; (h) UV–Vis absorption spectrum; (i) H2 evolution amount in visible light; (j) Gas adsorption detected by the QCM sensor. Adapted with permission from [18]. Copyright Wiley-VCH Verlag GmbH and Co. KGaA, Weinheim, 2017.
Nanomaterials 13 00499 g004
Figure 5. RN-g-C3N5: (a) SEM image; (b) TEM image; (c) UV–vis DRS spectra; (d) Tauc’s plots; (e) The photocurrent response; (f) EIS spectra; (g) MB degradation; (h) Recyclability of g-RN-C3N5. Adapted with permission from [20]. Copyright American Chemical Society, 2019.
Figure 5. RN-g-C3N5: (a) SEM image; (b) TEM image; (c) UV–vis DRS spectra; (d) Tauc’s plots; (e) The photocurrent response; (f) EIS spectra; (g) MB degradation; (h) Recyclability of g-RN-C3N5. Adapted with permission from [20]. Copyright American Chemical Society, 2019.
Nanomaterials 13 00499 g005
Figure 6. (a) Synthesis route; (b) molecular orbital overlap representation; (c) HR−XPS spectra (left: C1s, right: N1s); (d) CPMAS NMR spectra (left: 13C, right: 15N); (e) UV−vis spectra (blue: g−C3N4; red: g−C3N5); (f) steady state PL spectra; (g) PL lifetime decay curves; (h) X−band EPR spectra; Adapted with permission from [4]. Copyright American Chemical Society, 2019.
Figure 6. (a) Synthesis route; (b) molecular orbital overlap representation; (c) HR−XPS spectra (left: C1s, right: N1s); (d) CPMAS NMR spectra (left: 13C, right: 15N); (e) UV−vis spectra (blue: g−C3N4; red: g−C3N5); (f) steady state PL spectra; (g) PL lifetime decay curves; (h) X−band EPR spectra; Adapted with permission from [4]. Copyright American Chemical Society, 2019.
Nanomaterials 13 00499 g006
Figure 7. Co-C3N5 (a) Doping structure; (b) HR-TEM image; (c) HAADF-STEM images; (df) elemental mapping; (g) XANES spectra; (h) R-space of XANES; (i) FT-IR spectra; (j) XPS spectra of Co2p; (k) Tauc’s plot. (ah) Adapted with permission from [32]. Copyright Royal Society of Chemistry, 2022. (ik) Adapted with permission from [33]. Copyright Elsevier B.V., 2022.
Figure 7. Co-C3N5 (a) Doping structure; (b) HR-TEM image; (c) HAADF-STEM images; (df) elemental mapping; (g) XANES spectra; (h) R-space of XANES; (i) FT-IR spectra; (j) XPS spectra of Co2p; (k) Tauc’s plot. (ah) Adapted with permission from [32]. Copyright Royal Society of Chemistry, 2022. (ik) Adapted with permission from [33]. Copyright Elsevier B.V., 2022.
Nanomaterials 13 00499 g007
Figure 8. (a) Doping structure of P-C3N5; (b) K-edge XAFS spectra of 6P-gCN, and 10P-gCN samples and phosphorus reference compounds; (c) FT of the k2-weighted EXAFS spectrum (purple line) and the curve fitting result for the 10P-gCN sample (dotted line). Inset shows the k2-weighted EXAFS spectra; (d) UV-vis; (e) PL spectra; (f) A schematic illustration of B-C3N5; (g) ESR spectra; (h) Raman spectra; (i) UV-vis DRS spectra; and (j) the transient photocurrent response. (ae) Adapted with permission from [36]. Copyright American Chemical Society, 2021. (fj) Adapted with permission from [37]. Copyright Elsevier B.V., 2022.
Figure 8. (a) Doping structure of P-C3N5; (b) K-edge XAFS spectra of 6P-gCN, and 10P-gCN samples and phosphorus reference compounds; (c) FT of the k2-weighted EXAFS spectrum (purple line) and the curve fitting result for the 10P-gCN sample (dotted line). Inset shows the k2-weighted EXAFS spectra; (d) UV-vis; (e) PL spectra; (f) A schematic illustration of B-C3N5; (g) ESR spectra; (h) Raman spectra; (i) UV-vis DRS spectra; and (j) the transient photocurrent response. (ae) Adapted with permission from [36]. Copyright American Chemical Society, 2021. (fj) Adapted with permission from [37]. Copyright Elsevier B.V., 2022.
Nanomaterials 13 00499 g008
Figure 9. (a) Structures of g-C3N5 and NV-g-C3N5; (b) FT-IR spectra; (c) ESR spectra; (d) UV-vis spectra; (e) Tauc’s Plots; (f) EIS spectra; (g) FL spectra of g-C3N5 and NV-g-C3N5 (0.01, 0.02, 0.05, 0.1, and 0.2 represent the amounts of NaOH for vacancy production). (bf) Adapted with permission from [41]. Copyright American Chemical Society, 2020.
Figure 9. (a) Structures of g-C3N5 and NV-g-C3N5; (b) FT-IR spectra; (c) ESR spectra; (d) UV-vis spectra; (e) Tauc’s Plots; (f) EIS spectra; (g) FL spectra of g-C3N5 and NV-g-C3N5 (0.01, 0.02, 0.05, 0.1, and 0.2 represent the amounts of NaOH for vacancy production). (bf) Adapted with permission from [41]. Copyright American Chemical Society, 2020.
Nanomaterials 13 00499 g009
Figure 10. Schematic illustration of photoinduced carrier transfer on different heterojunctions: (a) type-I heterostructure; (b) type-II heterostructure; (c) Z-scheme heterostructure or S-scheme heterostructure; (d) Schottky junction.
Figure 10. Schematic illustration of photoinduced carrier transfer on different heterojunctions: (a) type-I heterostructure; (b) type-II heterostructure; (c) Z-scheme heterostructure or S-scheme heterostructure; (d) Schottky junction.
Nanomaterials 13 00499 g010
Figure 11. (a) CdS/C3N5 for RhB degradation; (b) CdS/C3N5 for H2 evolution; (c) g-C3N4/g-C3N5 for RhB and TC-HCl. (a) Adapted with permission from [53]. Copyright American Chemical Society, 2021. (b) Adapted with permission from [54]. Copyright Informa UK Limited, trading as Taylor and Francis Group, 2022. (c) Adapted with permission from [55]. Copyright Elsevier B.V., 2022.
Figure 11. (a) CdS/C3N5 for RhB degradation; (b) CdS/C3N5 for H2 evolution; (c) g-C3N4/g-C3N5 for RhB and TC-HCl. (a) Adapted with permission from [53]. Copyright American Chemical Society, 2021. (b) Adapted with permission from [54]. Copyright Informa UK Limited, trading as Taylor and Francis Group, 2022. (c) Adapted with permission from [55]. Copyright Elsevier B.V., 2022.
Nanomaterials 13 00499 g011
Figure 12. (a) g-C3N5/Bi4O5Br2 preparation routes; (b) Photocatalytic degradation of STZ by g-C3N5/Bi4O5Br2; (c) Photoreaction mechanism of STZ by g-C3N5/Bi4O5Br2; (d) Diagram of the g-C3N5/GO preparation; (e) U(VI) removal ratio; (f) Mechanism schematic of U (VI) extraction by C3N5/GO. (ac) Adapted with permission from [42]. Copyright Elsevier B.V., 2021 (df) Adapted with permission from [56]. Copyright Elsevier B.V., 2021.
Figure 12. (a) g-C3N5/Bi4O5Br2 preparation routes; (b) Photocatalytic degradation of STZ by g-C3N5/Bi4O5Br2; (c) Photoreaction mechanism of STZ by g-C3N5/Bi4O5Br2; (d) Diagram of the g-C3N5/GO preparation; (e) U(VI) removal ratio; (f) Mechanism schematic of U (VI) extraction by C3N5/GO. (ac) Adapted with permission from [42]. Copyright Elsevier B.V., 2021 (df) Adapted with permission from [56]. Copyright Elsevier B.V., 2021.
Nanomaterials 13 00499 g012
Figure 13. NH2-UiO-66/N-CN Z-scheme Heterojunction: (a) synthetic routes; (b) photocatalytic H2 production amounts; (c) charge density differences and energy-level diagrams; (d) the proposed Z-scheme charge transfer mechanism. Adapted with permission from [52]. Copyright American Chemical Society, 2022.
Figure 13. NH2-UiO-66/N-CN Z-scheme Heterojunction: (a) synthetic routes; (b) photocatalytic H2 production amounts; (c) charge density differences and energy-level diagrams; (d) the proposed Z-scheme charge transfer mechanism. Adapted with permission from [52]. Copyright American Chemical Society, 2022.
Nanomaterials 13 00499 g013
Figure 14. g-C3N5/Bi2MoO6 S-scheme heterojunction: (a) synthetic routes; (b) photocatalytic degradation of TC; (c) S-scheme charge transfer mechanism. Adapted with permission from [45]. Copyright Elsevier Inc, 2022.
Figure 14. g-C3N5/Bi2MoO6 S-scheme heterojunction: (a) synthetic routes; (b) photocatalytic degradation of TC; (c) S-scheme charge transfer mechanism. Adapted with permission from [45]. Copyright Elsevier Inc, 2022.
Nanomaterials 13 00499 g014
Figure 15. (a) Schematic illustration of HCNs@g-C3N5 synthesis; (b) The core-shell structure and proposed charge transfer pathway in HCNs@g-C3N5; (c) Schematic illustration of C3N5/RGO preparation; (d) The U(VI) removal ratio by C3N5/RGO under different conditions; (e) Mechanism schematic of U(VI) extraction by 3D C3N5/RGO. (ab)Adapted with permission from [61]. Copyright Elsevier B.V., 2021. (ce) Adapted with permission from [62]. Copyright Elsevier B.V., 2021.
Figure 15. (a) Schematic illustration of HCNs@g-C3N5 synthesis; (b) The core-shell structure and proposed charge transfer pathway in HCNs@g-C3N5; (c) Schematic illustration of C3N5/RGO preparation; (d) The U(VI) removal ratio by C3N5/RGO under different conditions; (e) Mechanism schematic of U(VI) extraction by 3D C3N5/RGO. (ab)Adapted with permission from [61]. Copyright Elsevier B.V., 2021. (ce) Adapted with permission from [62]. Copyright Elsevier B.V., 2021.
Nanomaterials 13 00499 g015
Figure 16. Adsorption and conversions of CO2: (a) Adsorption performance; (b) Adsorption selectivity for the binary mixtures CO2/H2 (20/80) and CH4/H2 (50/50); (c) CO2 adsorption on MCN-8E-T; (d,e) Calculated free energy diagram corresponding to the optimal path of the CO2 conversion to CH4 and CH3CH2OH on the g-C3N5, respectively; (f) Free energy diagram of HER (side reaction) on the g-C3N5; (g) and (h) CH4 and H2 evolution rate in photocatalytic CO2 reduction by g-C3N5: K0.01, respectively; (i) Schematic representation of the photocatalytic CO2 reduction using g-C3N5: K0.01. (ab) Adapted with permission from [29]. Copyright Royal Society of Chemistry, 2018. (c) Adapted with permission from [19]. Copyright Wiley-VCH Verlag GmbH and Co. KGaA, Weinheim, 2017. (df) Adapted with permission from [69]. Copyright Elsevier Inc, 2020. (gi) Adapted with permission from [35]. Copyright American Chemical Society, 2022.
Figure 16. Adsorption and conversions of CO2: (a) Adsorption performance; (b) Adsorption selectivity for the binary mixtures CO2/H2 (20/80) and CH4/H2 (50/50); (c) CO2 adsorption on MCN-8E-T; (d,e) Calculated free energy diagram corresponding to the optimal path of the CO2 conversion to CH4 and CH3CH2OH on the g-C3N5, respectively; (f) Free energy diagram of HER (side reaction) on the g-C3N5; (g) and (h) CH4 and H2 evolution rate in photocatalytic CO2 reduction by g-C3N5: K0.01, respectively; (i) Schematic representation of the photocatalytic CO2 reduction using g-C3N5: K0.01. (ab) Adapted with permission from [29]. Copyright Royal Society of Chemistry, 2018. (c) Adapted with permission from [19]. Copyright Wiley-VCH Verlag GmbH and Co. KGaA, Weinheim, 2017. (df) Adapted with permission from [69]. Copyright Elsevier Inc, 2020. (gi) Adapted with permission from [35]. Copyright American Chemical Society, 2022.
Nanomaterials 13 00499 g016
Figure 17. (ad) Initial structure (red circles: the binding position); (eh) corresponding optimized structures; (il) corresponding N−N bond distances; (m) Photocatalytic ammonia production rates of g-C3N5/NiCr-LDH; (n) Schematic illustration of a mechanism for NH3 production using g-C3N5/NiCr-LDH; (op) In-situ DRIFTS spectra of B-C3N5; (q) NH3 yields for g-C3N5 and B-C3N5 x (x = 1, 2, 3). (an) Adapted with permission from [70]. Copyright Elsevier B.V., 2022. (oq) Adapted with permission from [37]. Copyright American Chemical, 2020.
Figure 17. (ad) Initial structure (red circles: the binding position); (eh) corresponding optimized structures; (il) corresponding N−N bond distances; (m) Photocatalytic ammonia production rates of g-C3N5/NiCr-LDH; (n) Schematic illustration of a mechanism for NH3 production using g-C3N5/NiCr-LDH; (op) In-situ DRIFTS spectra of B-C3N5; (q) NH3 yields for g-C3N5 and B-C3N5 x (x = 1, 2, 3). (an) Adapted with permission from [70]. Copyright Elsevier B.V., 2022. (oq) Adapted with permission from [37]. Copyright American Chemical, 2020.
Nanomaterials 13 00499 g017
Figure 18. (a) O2-TPD spectra of C3N5 and 0.1−Ni−C3N5; (b) In situ DRIFTS spectra of 0.1−Ni−C3N5 on light irradiation; (c) Ion chromatography curves of NaNO3 and filtrate of reacted 0.1−Ni−C3N5 catalyst suspension; (d) Photocatalytic NO removal ratio for C3N5 and X−Ni−C3N5 catalysts; (e) Photocatalytic NO removal performance of 0.1−Ni−C3N5 in the presence of a series of trapping agents; (f) Proposed mechanism for photocatalytic NO removal on Ni-C3N5 material. Adapted with permission from [35]. Copyright Wiley-VCH GmbH, 2022.
Figure 18. (a) O2-TPD spectra of C3N5 and 0.1−Ni−C3N5; (b) In situ DRIFTS spectra of 0.1−Ni−C3N5 on light irradiation; (c) Ion chromatography curves of NaNO3 and filtrate of reacted 0.1−Ni−C3N5 catalyst suspension; (d) Photocatalytic NO removal ratio for C3N5 and X−Ni−C3N5 catalysts; (e) Photocatalytic NO removal performance of 0.1−Ni−C3N5 in the presence of a series of trapping agents; (f) Proposed mechanism for photocatalytic NO removal on Ni-C3N5 material. Adapted with permission from [35]. Copyright Wiley-VCH GmbH, 2022.
Nanomaterials 13 00499 g018
Figure 19. Photo-degradation processes of pollutants by g-C3N5-based materials.
Figure 19. Photo-degradation processes of pollutants by g-C3N5-based materials.
Nanomaterials 13 00499 g019
Table 1. Summaries of g-C3N5 composite materials on photocatalytic H2 production.
Table 1. Summaries of g-C3N5 composite materials on photocatalytic H2 production.
CatalystReaction ConditionsH2 Evolution
(mmol·g−1·h−1)
Quantum EfficiencyRef.
g-C3N5(MCN-8)300 W Xe lamp (λ > 420 nm)2.67 [18]
Catalyst (1g L−1)
TEOA (10 vol%)
Pt cocatalyst
Ultrathin C3N5 nanosheets300W Xe lamp (λ > 420 nm)0.03 [30]
Catalyst (0.5 g L−1)
TEOA (10 vol%)
NixSy-C3N5300 W Xe lamp (λ > 420 nm)35.4437.0%
(at 420 nm)
[63]
Catalyst (1.5 g L−1)
TEOA (10 vol%)
Pt cocatalyst
S-Ni (OH)2-C3N5300 W Xe lamp (λ > 420 nm)32.2230.9%
(at 420 nm)
[64]
Catalyst (1.5 g L−1)
TEOA (10 vol%)
Pt cocatalyst
Pt-C3N5300 W Xe lamp (λ > 420 nm)28.9628.7%
(at 420 nm)
[59]
Catalyst (1.5 g L−1)
TEOA (10 vol%)
CD/MoS2/C3N5300 W Xe lamp (λ > 420 nm)0.45 [65]
Catalyst (0.5 g L−1)
Na2S and Na2SO3 (0.35 M)
CdS/C3N5300 W Xe lamp (λ > 420 nm)0.50 [54]
Catalyst (0.5 g L−1)
TEOA (10 vol%)
Pt cocatalyst (3 wt%)
CdS/C3N5 (CCN)300 W Xe lamp (λ > 420 nm)2.69 [66]
Catalyst (0.5 g L−1)
Na2S and Na2SO3 mixtures (0.35 M)
Pt cocatalyst (3 wt%)
C3N5/Zn0.5Cd0.5S300W Xe lamp(λ > 420 nm)142.833.7%
(at 420 nm)
[67]
Catalyst (0.05 g L−1)
Na2S (0.35 M) and Na2SO3 (0.25 M)
NH2-UiO-66/N-CN-2visible light (λ ≥ 420 nm)3.946.8%
(at 420 nm)
[52]
Catalyst (0.05 g L−1)
TEOA (10 vol%)
Pt cocatalyst (2 wt%)
C3N4/rGO/C3N5300 W Xe-lamp(λ ≥ 400 nm)6.383.5%
(at 420 nm)
[68]
Catalyst (1 g L−1)
TEOA (10 vol%)
Pt cocatalyst (1.0 wt%)
Table 2. Summary of applications in photocatalytic NO removal using g-C3N5-based materials.
Table 2. Summary of applications in photocatalytic NO removal using g-C3N5-based materials.
CatalystsReaction ConditionsRemoval Efficiency (%)Reactive SpeciesRef.
NixSy-C3N5flow reactor and visible LED40%h+, ·OH, ·O2, and 1O2[63]
S-Ni (OH)2-C3N525 mg catalyst
NO (~600ppb at 1000 mL min−1)
A 30 W visible LED
42%h+, ·OH, ·O2, and 1O2[64]
TiO2(P25)-C3N5300 W Xe lamp (λ > 400 nm)
20 mg catalyst;
NO: 450 ppb and RH: 15%
67 %e, h+, and ·O2[72]
Ni-C3N5NO:600 ppb54%·OH, ·O2, and 1O2[60]
Table 3. Summary of the application of g-C3N5-based materials in degrading pollution.
Table 3. Summary of the application of g-C3N5-based materials in degrading pollution.
CatalystPollutantsReaction ConditionsReactive SpeciesRemoval EfficiencyRef.
RN-g-C3N5MB50 W halogen tungsten lamp·O2, ·OH98%, 120 min[20]
Catalyst (1 g L−1)
MB (20 mL, 20.0 mg L−1)
Ultrathin C3N5 nanosheetsMB300 W Xe lamp (400 nm)·O2, h+95%, 40 min[30]
Catalyst (0.5 g L−1)
MB solution (40 mL, 2.5 mg L−1)
Nv g-C3N5-0.1MB50 W halogen tungsten lamp·O2, ·OH95%, 120 min[25]
RhBCatalyst (1 g L−1)97%, 120 min
MOPollutant (20 mL, 40.0 mg L−1)95%, 120 min
CDs/MoS2/C3N5MB300 W Xe lamp (λ > 420 nm)1O2, ·O2, ·OH94%, 120 min[65]
Catalyst (0.02 g L−1)
MB (50 mL, 30 mg L−1)
CdS-MHPRhBSolar simulator (100 mW/cm2)·O2, ·OH, HO2, and e77%, 20 min
90%, 80 min
[53]
Catalyst (0.1 g L−1)
RhB (50 mL, 0.01mM)
g-C3N5/g-C3N4RhB
TC-HCl
300 W Xe lamp (λ > 420 nm)1O2, ·O2, and ·OH98%, 30 min
92%, 60 min
[55]
Catalyst (0.4 g L−1)
RhB (10 mg L−1, 50 mL)
TC-HCl (10 mg/L, 50 mL)
g-C3N5/MIL-101(Fe)/PANCMACarbamazepine300 W Xe lamp (λ >420 nm)h+, ·O2, and ·OH94%, 40 min[74]
ciprofloxacinCatalyst (0.1 g L−1)97%, 40 min
tetracyclineCarbamazepine (50 mL, 200 ng mL−1)98%, 40 min
AgCl/g-C3N5RhBA halogen lamp (300W)O2, h+96%, 30 min[49]
Catalyst (1 g L−1)
RhB solution (50 mL, 10 mg L−1)
Er3+/Tb3+@BiOBr-g-C3N5sulfamethoxazole500 W tungsten halogen lamp·O2, ·OH94%, 60 min[75]
Catalyst (1.3 g L−1)
SMX (75 mL, 10 ppm)
g-C3N5/Bi4O5Br2sulfathiazole (STZ)300 W Xe lamp·O2, h+, and ·OH100%, 60 min[42]
Catalyst (0.5 g L−1)
STZ (200 mL, 10 mg L−1)
Ag3PO4/C3N5TCHA 300 W Xe lamp (λ > 400 nm)·O2 and ·OH91%, 60 min[47]
Photocatalyst (1 g L−1)
TCH (50 mL, 20 mg L−1)
C3N5/Ag2CO3MB
TC-HCl
300 W xenon lamp, λ > 400 nm·O2 and h+97%, 90 min
98%, 100 min
[48]
Catalyst (1 g L−1)
MB (60 mg L−1, 50 mL, and pH = 8.0)
TC-HCl (50 mg L−1, 50 mL, and pH = 4.8)
Bi2WO6/g-C3N5Tetracycline
2-ercaptobenzothiazol
chlorpyrifos
under visible light (λ > 400 nm)h+, ·O2, and ·OH93%, 90 min[16]
Catalyst (0.6 g L−1)97%, 90 min
Pollutant (10 mg L−1, 50 mL)94%, 90 min
FeOCl/g-C3N5TC500 W Xe lamp (λ > 420 nm)500 W Xe lamp (λ > 420 nm)95%, 40 min[50]
Catalyst (1 mg mL−1)
TC (75 mL, 10 mg L−1)
H2O2 solution (30%, 200 μL)
CeTi2O6/g-C3N52,4 dichlorophenol300 W xenon lamp (λ > 420 nm)·O2, ·OH96%, 120 min[46]
Photocatalyst (1.6 g L−1)
2,4-DCP solution (75 mL, 10 ppm)
C3N5@NH2-MIL-125RhB300 W xenon lamp (λ > 420 nm)·O2, h+, and ·OH93%, 120 min[76]
Catalyst (0.5 g L−1)
RhB (100 mL, 10 mg L−1)
2D/0D
C3N5/ Bi2WO6
TC300 W Xe lamp (λ > 420 nm)·O2, h+, and ·OH94%, 60 min[44]
Catalyst (0.2 g L−1)
TC (20 mg L−1, 100 mL, and pH 5.2)
2D/0D
Bi2MoO6/C3N5
TC300 W Xe lamp (λ > 420 nm)·O2, ·OH, and h+88%, 75 min[45]
Catalyst (0.3 g L−1)
TC (20 mg L−1, 100 mL, and pH 5.2)
2D/2D
Bi2WO6@g-C3N5
TC300 W Xe lamp (λ > 420 nm)·O2, ·OH, and h+100%, 60 min[77]
Catalysts (0.4 g L−1)
TC (10 mg L−1, 50 mL)
2D/2D
Bi4O5Br2/g-C3N5
Ciprofloxacin
bisphenol-A
Xe lamp at 500 W (165 mW/cm2)·O2, ·OH, and h+94%, 60 min
92%, 80 min
[78]
Catalysts (0.67 g L−1)
Pollutant (20 mg L−1, 75 mL)
2D/0D
C3N5/ Bi2WO6
Cr (VI)300 W Xe lamp (λ > 420 nm)·O2 and e97%, 50 min[44]
Catalyst (0.2 g L−1)
Cr (VI) (10 mg L−1, 100 mL, and pH 2.5)
2D/0D
Bi2MoO6/C3N5
Cr (VI)300 W Xe lamp (λ > 420 nm)·O2 and e97%, 60 min[45]
Catalyst (30 mg)
Cr (VI) (10 mg L−1, 100 mL, and pH 2.5)
2D/2D
C3N5/GO
U(VI)300 W Xe lamp (λ > 420 nm)e96%, 90 min[56]
Catalyst (0.5 g L−1)
U(VI) solution (10 ppm, 100 mL)
2D/2D
C3N5/RGO
U(VI)300 W Xe lamp (λ > 420 nm)e95%, 100 min[62]
Catalyst (0.2 g L−1)
U(VI) (10 mg L−1, 100 mL)
pH 5.0
T = 298 K
Bi4O5I2/g-C3N5Hg06 W LED lamp (λ > 400 nm)·O2 and h+93%, 60 min[43]
Catalyst (40 mg)
Mercury vapors (65 μg m−3)
gas flow rate of 1.2 L min−1
C3N5SMX300 W xenon lamp (λ > 420 nm)·O2, h+, SO4⋅, 1O2, and ·OHPMS/C3N5/Vis[28]
Catalyst (0.5 g L−1)67%, 60 min
PMS (0.125 g L−1)PDS/C3N5/Vis
SMX (5 mg L−1)70%, 60 min
PDA-g-CN-1.0SMXCatalyst (50 mg L−1)C−PMS * complexes100%, 20 min[39]
PMS (1mM)
SMX (10 mg L−1)
Co-C3N5PCB28Catalyst (0.2 g L−1)SO4· and ·OH96%, 30 min[32]
PMS (2.0 mM)
PCB28 (0.5 mg L−1)
Co-C3N5SMX500 W Xe lamp (λ > 420 nm)high-valent cobalt oxide (Co (IV)) species100%, 20 min[33]
Catalyst (0.5 g L−1)
PMS (1.0 mM)
SMX solution (30 mL, 10 mg L−1)
U-C3N5SMZvisible light (λ > 420 nm)·OH, ·O2, 1O2, ·CO3, and h+94%, 120 min[79]
Catalyst (0.4 g L−1)
SPC (0.1 g L−1)
SMZ (100 mL, 10 mg L−1)
* indicates that C and PMS form a complex (C-PMS*)
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

Liu, J.; Wang, S.; Zhao, C.; Zheng, J. Engineered g-C3N5-Based Nanomaterials for Photocatalytic Energy Conversion and Environmental Remediation. Nanomaterials 2023, 13, 499. https://doi.org/10.3390/nano13030499

AMA Style

Liu J, Wang S, Zhao C, Zheng J. Engineered g-C3N5-Based Nanomaterials for Photocatalytic Energy Conversion and Environmental Remediation. Nanomaterials. 2023; 13(3):499. https://doi.org/10.3390/nano13030499

Chicago/Turabian Style

Liu, Juanjuan, Shuaijun Wang, Chaocheng Zhao, and Jingtang Zheng. 2023. "Engineered g-C3N5-Based Nanomaterials for Photocatalytic Energy Conversion and Environmental Remediation" Nanomaterials 13, no. 3: 499. https://doi.org/10.3390/nano13030499

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