Next Article in Journal
Performance of TiO2/UV-LED-Based Processes for Degradation of Pharmaceuticals: Effect of Matrix Composition and Process Variables
Next Article in Special Issue
Slow-Relaxation Behavior of a Mononuclear Co(II) Complex Featuring Long Axial Co-O Bond
Previous Article in Journal
Effects of Fe, Mn Individual Doping and (Fe, Mn) Co-Doping on Ferromagnetic Properties of Co2Si Powders
Previous Article in Special Issue
Muon Irradiation of ZnO Rods: Superparamagnetic Nature Induced by Defects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Comprehensive Review of Graphitic Carbon Nitride (g-C3N4)–Metal Oxide-Based Nanocomposites: Potential for Photocatalysis and Sensing

by
Amirhossein Alaghmandfard
and
Khashayar Ghandi
*
Department of Chemistry, University of Guelph, Guelph, ON N1G 2W1, Canada
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(2), 294; https://doi.org/10.3390/nano12020294
Submission received: 10 December 2021 / Revised: 27 December 2021 / Accepted: 5 January 2022 / Published: 17 January 2022
(This article belongs to the Special Issue Advances in Nanostructured Semiconductors and Heterojunctions)

Abstract

:
g-C3N4 has drawn lots of attention due to its photocatalytic activity, low-cost and facile synthesis, and interesting layered structure. However, to improve some of the properties of g-C3N4, such as photochemical stability, electrical band structure, and to decrease charge recombination rate, and towards effective light-harvesting, g-C3N4–metal oxide-based heterojunctions have been introduced. In this review, we initially discussed the preparation, modification, and physical properties of the g-C3N4 and then, we discussed the combination of g-C3N4 with various metal oxides such as TiO2, ZnO, FeO, Fe2O3, Fe3O4, WO3, SnO, SnO2, etc. We summarized some of their characteristic properties of these heterojunctions, their optical features, photocatalytic performance, and electrical band edge positions. This review covers recent advances, including applications in water splitting, CO2 reduction, and photodegradation of organic pollutants, sensors, bacterial disinfection, and supercapacitors. We show that metal oxides can improve the efficiency of the bare g-C3N4 to make the composites suitable for a wide range of applications. Finally, this review provides some perspectives, limitations, and challenges in investigation of g-C3N4–metal-oxide-based heterojunctions.

Graphical Abstract

1. Introduction

Graphitic carbon nitride (g-C3N4) is a polymeric, visible-light-active photocatalyst with a bandgap of ~2.7 eV (~460 nm), that was introduced since 2006 [1]. g-C3N4 has become an important material in chemistry, physics and engineering because of its facile, low-cost, environmentally-friendly preparation methods with promising stability and good physicochemical properties for use in a wide range of applications [2]. Compared with other semiconductors, g-C3N4 can be easily synthesized by various methods with desirable electrical structures as well as morphologies, and high thermal stability up to 600 °C in the air [3,4].
The most common precursors used to prepare g-C3N4 are melamine, dicyandiamide, cyanamide, urea, thiourea, and ammonium thiocyanate. Among different types of carbon nitrides such as α-C3N4, β-C3N4, cubic C3N4, pseudocubic C3N4, with bandgaps of around 5.49 eV, 4.85 eV, 4.30 eV, and 4.13 eV, respectively, g-C3N4 is the most stable phase under ambient conditions [2]. In order to enhance the performance and modulate the properties of g-C3N4, researchers have proposed different methods such as doping and making heterojunction with other materials. Examples of these materials are metal oxides, metal sulfides, noble metals, and carbonaceous nanomaterials [5,6,7,8,9,10,11,12]. Among them, metal oxides are the most common ones to improve the efficiency of g-C3N4, e.g., increasing the light absorption and reducing the recombination of electrons and holes by promoting the separation of charge carriers. This is mainly due to their suitable band structures [13,14,15,16,17,18].
The g-C3N4 structure has been widely used in many applications, in particular in energy-related applications. Energy consumption to provide electricity and heat will rise to twice its current consumption by 2050, which is mainly due to industrialization, urbanization, and population growth [19,20]. The consumption of fossil fuels, such as natural gas, coal, and oil, should be decreased as their usage results in detrimental environmental impacts [2,19,20]. Two remedies are solar energy and photocatalysis [21,22,23,24]. Both require suitable semiconductors such as g-C3N4 with superior activities for different catalytic reactions, such as organic pollutants degradation, H2, and O2 generation by water splitting and CO2 reduction to hydrocarbon fuels [14,25,26,27,28,29]. The g-C3N4 can also be used for water disinfection and bacterial control [25,30].
It is evident from Figure 1 that the number of publications, collected from the Scopus database, has been growing fast from 2012 in the field of g-C3N4 and g-C3N4–metal oxide-based heterojunctions. Figure 1a shows the number of all publications on g-C3N4 since 2012, showing that this topic is among the hot research areas. It is therefore important to provide a comprehensive review of the g-C3N4–metal oxide composites. To the best of our knowledge, no publication reviews a wide range of papers to explain the applications and structure of these heterojunctions comprehensively.
This review covers the research up to 1 October 2021. We highlighted some general information about the structure and characterization of the bare g-C3N4. We also discussed some modifications such as doping to improve the g-C3N4 properties. As well, we have also summarized the research on modification of the structure and properties, to enhance the efficiency of g-C3N4 for different applications, via combining g-C3N4 with metal oxides such as TiO2, ZnO, iron oxide, WO3, and tin oxide. After reviewing g-C3N4–metal oxides, we focus on the applications of these kinds of heterojunctions. In the last part of this review article, we suggest some potential investigations for the future in this field that have not been conducted to this date to our knowledge. We also recommend some studies to better understand the nature of these heterojunctions. Most graphs in this review are reproduced by us from the data in original research papers cited, unless stated otherwise in the captions where the permission has been obtained.

2. Structure and Properties of g-C3N4

Melamine, melam, melem, and melon are recognized as heptazine- and triazine-based molecular compounds (coplanar tri-s-triazine unit as the elementary structural motif of g-C3N4 structure) to prepare g-C3N4. As illustrated in Figure 2, triazine (C3N3) and tri-s-triazine/heptazine (C6N7) rings are the basic tectonic units of g-C3N4.
Yan et al. properly studied the phase transition during heating of melamine from room temperature to 1000 °C with a heating rate of 10 °C/min [31]. Figure 3a shows that melamine sublimation and thermal condensation occur at 297 to 390 °C, observed from significant endothermic differential scanning calorimetry (DSC) peak and drastic thermogravimetric analysis (TGA) weight loss. Other endothermic peaks at 545 and 630 °C are attributed to the materials’ deamination and decomposition, respectively.
Figure 3b shows the g-C3N4 thermal stability and phase transitions in an open system. This figure shows that g-C3N4 has high stability below 600 °C, which is 30 °C lower than the melamine decomposition temperature (630 °C). Beyond this temperature, g-C3N4 starts to decompose into small molecules (e.g., CO2 and NH3) [31]. The stability of g-C3N4 is higher in the semiclosed ammonia atmosphere than in the open system, mainly due to the inhibition of deamination under ammonia atmosphere. The TGA result illustrated that there is no residue at 750 °C [31]. The annealing temperature is a vital factor in the preparation and properties of the final g-C3N4 structure. Praus et al. used melamine as a precursor to synthesize g-C3N4 in an air atmosphere [32]. The TGA results, illustrated in Figure 3c, demonstrates that by heating the melamine to 400 °C melen is formed. The main reason for the weight loss at this temperature range is the elimination of ammonium. By further heating up to 600 °C, melon is obtained by melem polymerization [33]. In other words, at higher temperatures, the g-C6N9H2 or g-C3N4.5H is more stable than g-C3N4. Figure 3d illustrates that the C/N molar ratio is temperature dependent. The main difference in the C/N value of the synthesized g-C3N4 compared to the theoretical value of 0.75 for g-C3N4 is due to the incomplete condensation of the amino groups of melon and the low degree of polymerization. Reaction atmosphere is another factor affecting the g-C3N4 by providing defects and disordered structures. For instance, in the H2 atmosphere, by dicyanamide thermal condensation, more nitrogen vacancies were formed [34].

3. g-C3N4 Characterizations

The presence of X-ray diffraction (XRD) peaks at about 13 and 27° is an indication of the formation of g-C3N4, corresponding to the (002) and (100) diffraction patterns, respectively [35,36,37]. Paul et al. showed the effects of calcination temperature on the naked g-C3N4 using XRD data, Fourier transform infrared spectra (FT-IR), bandgap structure, Brunauer–Emmett–Teller (BET), and photocatalytic recyclability data [36]. Due to the polycondensation of melamine at lower than 400 °C, the crystallinity of the g-C3N4 structure is detected (Figure 4a). In contrast, Figure 4b demonstrates that no significant differences are revealed in FTIR spectra at different calcination temperatures. In the FTIR spectra, the characteristic bands with high intensity at the wavenumber about 1640, 1569, 1412, 1328, and 1240 cm−1 are related to the stretching modes of C=N and C–N heterocycles. Besides, the strong peaks at 815 cm−1 are attributed to the s-triazine units. The broad range peak between 3000 cm−1 and 3500 cm−1, is due to the N-H stretching and the remaining water molecules in the structure [35,36,37,38,39]. The BET specific surface areas were estimated to be 37.8, 73.7 and 65.6 m2·g−1 for g-C3N4 synthesized at the calcination temperature of 450, 550 and 650 °C, respectively. Calcination temperature also affects the g-C3N4 bandgap. The bandgap when calcinated at 550 °C is narrower than when calcinated at 450 and 650 °C (Figure 4c) [36,38,40]. When the temperature increases to 700 °C, a higher bandgap is observed. In other words, the higher the degree of g-C3N4 polymerization, the larger the π-plane conjugation degree of heptazine rings via N2 atoms, at the higher temperature.
To assess the catalytic activity, ultraviolet-visible (UV-vis) diffuse reflectance spectrums (DRS) of the materials were determined at different temperatures [36]. As discussed earlier the results show that the bandgap of g-C3N4 prepared at 550 °C is narrower than those formed at 450 and 650 °C, so the structure prepared at the 550 °C absorbs more visible light, which can lead to the more appealing photoactivity. Moreover, the adsorption and degradation efficiencies for methylene blue dye by g-C3N4 prepared at 500 °C is 34.4% and 62.6%, which is higher than those synthesized at 450, 550, 600, and 650 °C [36]. Figure 4d reveals that the prepared g-C3N4 at a calcination temperature of 550 °C also showed good stability even after four cyclic runs. The pH, and catalyst loading are also important factors in the adsorption and degradation efficiency [36].
Chen and coworkers illustrated that the PL depends on the condensation temperature (Figure 5a), showing the strong and stable emission in the range of 450–510 nm, mainly due to the π*–π, and π*–LP transitions [41]. The increase in the amount of the tri-s-triazine content at high temperature leads to the higher π states and cause orbital overlap and reduces the PL intensity. The UV-Vis spectra of the g-C3N4 at different synthesis temperatures are illustrated in Figure 5b. Like the PL emission peak, the strong absorption peaks at 450 nm and 500 nm are due to the π*–π and π*–LP transitions. Another study, conducted by Dias et al., concluded that during thermal treatment of g-C3N4, the enhanced porosity is mainly due to the creation of N vacancies and defects/holes within the nanosheets, resulting in the improvement of optoelectrical properties. Not only the treatment leads to the appearance of n → π∗ transitions, but we can also observe red shift in the absorption spectra. Moreover, it can enhance the photocatalytic activities by improving the carrier separation, reducing charge recombination, which is mainly due to the presence of trap states [42]. Furthermore, researchers also investigated the effect of the different dopants on the absorption properties of g-C3N4 [43]. These results showed the effect of temperature and dopants on the optical properties of g-C3N4. Yuan et al. showed the photograph of the melamine and g-C3N4 synthesized at different temperatures, graphically showing the color variation in the deionized water under UV light (365 nm) (Figure 5c) [44]. By increasing the synthesis temperature, the emitted color varied from blue–violet to green, and the intensity went through a maximum at 550 °C. The impact of the temperature on the PL spectra was demonstrated by the 5–10 nm blue-shift at lower temperatures, due to a less delocalized orbital resulting in a larger bandgap. Figure 5d also reveals strong PL emission spectra with a much smaller blue-shift. This research is among the rare studies investigating the effect of the environment on PL emission spectra. The g-C3N4 time-resolved PL spectra shows 5 ns electron-hole recombination at 25 °C [45].

4. g-C3N4 Preparation

Synthetic techniques for graphitic carbon nitride (g-C3N4) were first reviewed by Thomas et al. in 2008 [26]. g-C3N4 can be synthesized with various techniques such as chemical vapor deposition (CVD), solvothermal, and plasma sputtering deposition. Chemical vapor deposition to deposit g-C3N4 thin film on indium-tin-oxide (ITO), was performed by Ye et al. [27]. In this method, the mixture of thiourea and melamine was put at the bottom of a crucible with ITO substrate above it and finally transferred to the muffle furnace. CVD has several advantages and disadvantages. To be more specific, thin films prepared by CVD are cohesive in all dimensions, which is suitable for elaborately shaped pieces and helps users to fill the insides, undersides, high aspect ratio holes, etc. CVD does not require high vacuum and can deposit a wide variety of materials to prepare high purity composites. In contrast, physical vapor deposition (PVD) such as sputtering requires a high vacuum atmosphere. The drawback of CVD is that some CVD precursors are costly, and can be highly toxic, explosive, or corrosive, such as Ni(CO)4, B2H6, and SiCl4, respectively. The by-product of this method, including CO, or HF, can also be hazardous. The substrates are limited since they should tolerate high temperatures [46].
The graphitic carbon nitride nanocone arrays were grown onto the Ni-coated Si (100) substrate, using plasma sputtering deposition [47]. This method requires a vacuum chamber and a plasma source with a discharge cavity. Thermal condensation is an economical, energy-efficient approach, which has a higher chance for scaleup for commercialization. The solvothermal methods have some drawbacks such as more synthetic steps compared to the thermal condensation method; however, they can be cost-effective, low energy consumption methods with more controllable size and morphology [48,49,50].
Several C−N precursors such as urea, thiourea, melamine, cyanamide, and dicyanamide are used for g-C3N4 synthesis (Figure 6a). The procedure is depicted in Figure 6b [51,52,53,54,55]. Among them, cyanamide and its derivates such as dicyanamide suffer from low solubility and high cost [56]. Pham and Shin showed that urea and melamine cause poor interconnection of g-C3N4 to NiTiO3 since melamine provides a segregated g-C3N4 structure with no connection to the NiTiO3 phase, and urea makes a condensed g-C3N4 structure by releasing oxygen-containing gas during the thermal condensation [56]. Dicyandiamide and thiourea with higher reactivity towards a polymerization reaction create strong Ti−N bonds in the composite, and allow for charge carrier formation, which is important for degradation of organic contaminants [57]. The order of photodegradation ability of bisphenol A (BPA) is g-C3N4-Melamine in N2 atmosphere ≈ g-C3N4-dicyandiamide > g-C3N4-dicyandiamide in N2 atmosphere > g-C3N4-Melamine > g-C3N4-Urea ≈ g-C3N4-Urea in N2 atmosphere [58]. The differences between the catalytic activities are mainly due to the different preparation procedures, which can change the type, the density of active sites, the network of sp2 hybridized carbon, nitrogen, and oxygen-containing functional groups [58]. Jung et al. showed that the precursor could affect the morphology of the g-C3N4-based systems [59]. The good binding of dicyandiamide (DCDA) to ZnO nanoparticles, lead to the formation of the core-shell morphology DCDA-CNZ composite, resulting in improving the degradation of methylene blue by charge transfer [59]. In contrast, due to the weak interaction of thio and urea with ZnO, thio and urea-CNZ have a porous and segregated morphology and produce gases during the polymerization [59].
As shown in Figure 6b, melamine can form from urea, thiourea, and cyanamide (dicyanamide), and then converted to tri-s-triazine (and melam) rings at ~335 and ~390 °C, respectively, and subsequently, g-C3N4 discovered from tri-s-triazine polymerization, by heating to 520 °C. Besides, the skeleton becomes prepared over 600 °C, and then beyond 700 °C, g-C3N4 is entirely decomposed into small molecules (e.g., NH3). Figure 7a,b shows the urea and thiourea conversion mechanism into melamine. In this reaction, oxygen atoms in the urea structure help to facilitate g-C3N4 condensation and improve stability [60]. CO2 and NH3 are the by-products of the reaction of the melamine formation from urea, which can be recycled again to urea. In the formation of the g-C3N4, the presence of a crucible lid is so important since not only it prevents the gasses from escaping, but it also provides high pressure in the synthesis atmosphere, which is necessary for the preparation of g-C3N4. In other words, a covered alumina crucible should be used during thermal analysis to avoid melamine sublimation.
During the heating reaction, the created gas bubbles act as soft templates to produce 24 nm pores in the yellow-colored graphitic carbon nitride [61]. Thiourea is used as another precursor for the g-C3N4 formation, and sulfur content improves the connectivity and packing of g-C3N4 sheets [60]. The overall conversion of urea and thiourea to melamine is endothermic. The first endothermic reaction of changing urea and thiourea to melamine is at temperatures higher than their melting temperatures, which are ~133 and ~180 °C, respectively, while at the higher temperature, melamine and heptazine were prepared in low pressure at, e.g., atmospheric pressure. To be more specific, the second reaction requires preheating to 260–280 °C to decompose urea in the presence of ammonia, passed over the activated alumina, silica gel, silica-alumina gel, or alumina gel. In order to completely form melamine, the vapors obtained from the first reaction should be maintained at ~400 °C [62]. It should also be noted that the g-C3N4 can be prepared by cyanamide and dicyanamide. Specifically, by polycondensation of cyanamide molecules and dicyandiamide, melamine was prepared at ~203 and ~234 °C, respectively [2].

5. g-C3N4 Modifications

5.1. Doping

The presence of heptazine ring in the g-C3N4 affects electronic structure, toxicity, and density, and is important for applications, especially as biosensors, for photocatalytic hydrogen evolution, and CO2 conversion. The density-functional theory (DFT) showed that the bandgap of the fully condensed g-C3N4 is lower than melem, and polymeric melon, which are 2.1 eV, 3.5 eV, and 2.6 eV, respectively [63,64,65]. As previously reported, the polymeric melon has a bandgap close to the defect containing bulk g-C3N4 [64]. There are mainly two types of experimental results to calculate the semiconductor bandgap. Specifically, the position of the conductive band (CB) and valance band (VB) can be measured by electrochemical impedance spectra (EIS), and Mott–Schottky (M–S) curve [63]. The result of both methods showed that the position of the conduction band (CB) and valance band (VB) of the g-C3N4 is about −1.3 eV and 1.4 eV, respectively [3,66]. The semiconductors’ band edges can be tuned by functionalizing, doping, compositing with other materials. Liu et al. reviewed element-doped carbonized nitrogen in detail and investigated their organic pollutants degradation applications [3]. Ai et al. showed that the bandgap value of the phosphate doped g-C3N4 decreased from 2.57 eV to 2.49 eV, 2.43 eV, and 2.41 eV by increasing the content of the P element [67]. Other researchers demonstrated that the higher O, Na, Ag, and Co content doped to the g-C3N4 structure leads to the lower bandgap value [68,69,70,71]. Li et al. investigated the effect of different ratios of Sm to g-C3N4. They illustrated that by varying the percentile molar ratios of Sm(NO3)3·5H2O with melamine from zero to 0.01%, 0.025% and 0.05%, the bandgap decreased from 2.63 eV to 2.57 eV, 2.50 eV, and 2.44 eV, respectively, so Sm narrowed the g-C3N4 bandgap [72]. Table 1 is illustrated the effect of different dopants on the band edge position. We will further discuss and analyze the properties and applications of different g-C3N4–metal oxides. g-C3N4 is a metal-free semiconductor, which possesses a narrow bandgap suited for visible light absorption (45% of solar energy output) [66]. To have an in-depth investigation of the optical properties of the g-C3N4, we will investigate some characteristics of the synthesized g-C3N4 such as photoluminescence (PL) and UV-vis spectra. The origin and nature of the PL emission come from three different transition pathway including π*–π, σ*—the nitride atom bridge’s lone pair (LP), and the π*–LP transition.

5.2. Metal Oxide-Based g-C3N4 Nanocomposite

Different types of metal oxides, such as TiO2, ZnO, WO3, iron oxide, tin oxide, etc., can improve the photocatalytic efficiency of the g-C3N4 by reducing the electrons-holes recombination and promoting the charge carriers’ separation. Consequently, metal oxide-based g-C3N4 nanocomposites can be used in different applications with enhanced electric, magnetic, and photocatalytic properties, such as H2 generation, CO2 reduction, NO oxidation, degradation of organic and inorganic dyes and other organic material, removal of toxic metal species, especially Cr (VI) from water, antibodies decontamination, solar cells, sensing, etc. [100,101,102,103,104]. In this part of the review, some metal oxide-based g-C3N4 heterojunction structures are compared with the g-C3N4.
There are five types of charge carrier separation for g-C3N4–metal oxide photocatalysts:
(1)
Type I heterojunction,
(2)
Type II heterojunction,
(3)
Z-scheme heterojunction
(4)
p-n heterojunction,
(5)
Schottky junction.
Most g-C3N4–metal oxide photocatalysts show type II and Z-scheme mechanisms for charge carrier separation. In this section, we will discuss these two heterojunction types.
In type II heterojunctions, two semiconductors are bound to form a stable heterojunction, and the position of the VB of semiconductor A is higher than that of semiconductor B. In this case, because of the difference in voltages, the photoinduced hole migrated from the VB of semiconductor B to that of semiconductor A (Figure 8a). On the other side, electrons are transferred from CB of semiconductor A to that of semiconductor B. The enhanced electrons and holes separation will reduce the rate of the recombination and promote the electrons’ lifetime. The construction of type II systems is highly desired for photocatalysis for different applications [7].
The other heterojunction type is the direct Z-scheme photocatalytic system, which was initially suggested by Bard et al. in 1995 [105]. As shown in Figure 8b, in this heterojunction type, the generated electrons on the CB of semiconductor B transfer to the VB of semiconductor A and combines with the photogenerated holes. This type of photocatalysts can be helpful for both reducing the recombination by an increase of the electrons and holes separation and improving the redox ability [7]. Even if the electrons, and holes combine and generate hν in these heterojunctions, other photogenerated carriers can replace them.
Different patterns of the migration of electrons and holes are due to the driving force of the electric field, formed by the band edge positions of different semiconductors. Depending on the electrical field direction, charge carriers start moving to reduce the energy of the system. Thus, different systems are defined based on the migration of electrons and holes in the heterojunction. The difference in each type is mainly due to the movement of electrons and holes. All types can be used in different photocatalytic activities.

5.2.1. TiO2-g-C3N4

Among the investigated semiconductor photocatalysts, TiO2 has a suitable conduction band position, excellent stability, cost-effective preparation approach and is one of the most promising catalytic materials [106]. Fujishima and Honda were pioneers who researched using TiO2 photocatalytic behavior in 1972 [107]. To improve TiO2′s photocatalytic efficiency, researchers would like to reduce the bandgap of the system by doping with other elements and compositing with other compounds to absorb visible light energy [108]. To deal with TiO2′s limitations, researchers have been using doping elements and compositing with organic material, such as conjugated polymers and g-C3N4, since they have a narrow bandgap [108]. Boron is an effective dopant, which can improve the photocatalytic applications of TiO2 coupled with carbon nitride, as shown by Christoforidis and coworkers [109]. In another study conducted by this research group, TiO2 and carbon nitride nanosheets were synthesized by hydrothermal in-situ approach, improving the catalytic application. The mentioned materials have high porosity to ensure a high concentration of reactants in the vicinity of catalytic sites, used for CO2 reduction [110].
Various synthetic methods such as co-calcination, hydrothermal treatment, solvothermal, and microwave-assisted to prepare the g-C3N4-TiO2 heterojunction [111,112,113]. The preparation of g-C3N4-TiO2 heterojunction widely includes the hydrothermal and calcination method [114]. Figure 9 illustrates the schematic of the TiO2/g-C3N4 preparation route.
The conductive band of g-C3N4 (ECB = −1.4 eV) is higher than those of anatase TiO2 (ECB = −0.5 eV), so the proper bandgap alignment helps to reduce the electron-hole recombination by increasing their separation and boost the space charge accumulation at the interface. Recombination is defined as a process of electrons and holes annihilation. The most common type of recombination is known as radiative recombination, which occurs when electrons and holes in a conductive and valance band, respectively, recombine and emit a photon. As discussed in the following sections, the presence of electrons and holes will assist us in different stages of photocatalytic activities. So, suppressing the charge recombination is crucial in the photocatalytic applications of semiconductors and heterojunctions. At the end of section three, we talked about how g-C3N4-based heterojunction separates the position of the electrons and holes. The charge carrier separation makes electron-holes recombination less favorable, so we can use these heterojunctions for photocatalytic activities.
Figure 10a illustrates the conduction, valance, and bandgap position of the g-C3N4 and TiO2 and reveals that the g-C3N4-TiO2 structure improves the photo-induced electrons flow from the g-C3N4 conductive band to that of TiO2, which promotes the photoelectrical ability of the final composite. Thus, the photogenerated electrons tend to accumulate in the TiO2 conductive band since the conductive band of TiO2 is more positive than that of g-C3N4, (Figure 10a). In contrast, the holes transfer in an inverse way, which can provide type II heterojunction [112,115]. Due to the high recombination barrier, the heterojunction provides a high interfacial area for facilitating carrier transformation and separation to suppress the electron-hole recombination to improve the photocatalytic activity.
Kočí et al. successfully deposited TiO2 on the g-C3N4 surface by hydrothermal approach followed by calcination processing [116]. In this research, the mixture of TiO2 and g-C3N4, prepared by thermal hydrolysis and polycondensation from melamine, respectively, should be mixed in distilled water for 16 h in an air atmosphere dried at 60 °C. Finally, the dried sample should be maintained at 450 °C in a covered crucible with a heating ramp of 15 °C/min in the air in a muffle furnace. The highest photocatalytic performance was obtained under UVA (λ = 365 nm) irradiation compared to other types of ultraviolet (UV) rays (UVB, and UVC), which is attributed to the high separated charge carrier. The low rate of recombination leads to the great photocurrent stability. In another work on g-C3N4-TiO2 heterojunction structure, Alcudia-Ramos et al. demonstrated that the prepared heterojunction has a higher photocatalytic efficiency than the individual g-C3N4 and TiO2 [112]. The researcher also showed that solvothermal synthesis enhances tri-s-triazine’s carbonization [117]. Miranda et al. used an impregnation method to prepare g-C3N4-TiO2 [118]. The presence of the g-C3N4 in the heterojunction endows the high specific area to the structure. The researchers also used the photochemical reduction method to prepared g-C3N4-TiO2 based nanocomposite. They also demonstrated that in the TiO2/g-C3N4/G composite, TiO2 was a semiconductor to capture visible light and also prevents g-C3N4/G stacking [119].
Several parameters affect the g-C3N4-TiO2 heterojunction structure. Various research works have been investigated the effect of the g-C3N4′s ratio. Wang et al. prepared this microstructure with 10%, 30%, 50%, 70% of g-C3N4, which are labeled with (x = 10, 30, 50, and 70) [110]. All the XRD results from different publications have shown the same diffraction pattern. As a case in point, the XRD pattern of g-C3N4-TiO2 is illustrated in Figure 10b. In the TiO2 diffraction pattern, the peaks at 25.5, 37.7, 48.2 and 54.1 °C are related to the (101), (004), (200) and (105) planes, respectively. Furthermore, there is no g-C3N4 characteristic peak at 13.1 °C in the g-C3N4-TiO2 hybrid sample, which may be attributed to the low crystallinity of g-C3N4 compared to the TiO2 in the microstructure. However, the peak at 27.5 is obviously observed in all samples. The more g-C3N4 ratio in the structure leads to the higher XRD intensity at 27.5 °C and the lower peak intensity at 25.5 °C [120]. TiO2 possesses a high crystallinity so the characteristic peak at 25.5 °C is higher than 27.5 °C. There are no obvious changes in other characteristic peaks in the composite XRD patterns. The FT-IR spectra of the g-C3N4-TiO2 structures are demonstrated in Figure 10c [120]. In all FT-IR analyses, stretching vibration of the Ti–O and Ti–O–Ti is observed at about 475 cm−1. Additionally, peaks at 3419 cm−1 are attributed to the absorbed moisture and hydroxyl group in the structure. The peak at 2360 cm−1 that appeared in all samples is mainly due to the adsorbed CO2. All the characteristic peaks at TiO2 and g-C3N4 are observed in the composite samples. The pore size distribution of the g-C3N4, TiO2, g-C3N4-TiO2 is illustrated in Figure 10d [120]. The pore size at ~3.5 nm, 10 nm, and 3.8 nm, 32 nm is due to the presence of TiO2, and g-C3N4, respectively. Additionally, the BET specific area of the g-C3N4-TiO2 hybrid structure was increased at the higher g-C3N4 content. Finally, the higher the bandgap of TiO2, the lesser the absorption wavelength at higher than 400 nm (Figure 10e). The g-C3N4 revealed the absorption band extending to about 430 nm, which is mainly due to the low bandgap [120]. The blue shift in the absorption spectra was also demonstrated at a higher g-C3N4 ratio in the structure. It was shown bandgaps calculation of the g-C3N4, TiO2, g-C3N4-TiO2. The results revealed that the bandgap of CNT50 samples is 2.92 eV, which is like the bulk g-C3N4 with the bandgap of 2.9 eV and narrower than the bandgap of TiO2 (3.20 eV). Due to the narrow bandgap and heterojunction formation, the hybrid composite not only suggested the higher generation of the electron and hole but also improved photocatalytic activity [120,121]. Li et al. synthesized g-C3N4@TiO2 hollow sphere nanostructure with high crystallinity [122]. In this work, researchers used the different ratios of TiO2 hollow sphere and melamine (1:2, 1:4, and 1:8) in the solution, which are called HS-CNTO1, HS-CNTO2, and HS-CNTO3, respectively. They demonstrated that the recombination rate of the photogenerated electron-holes decreased by introducing the TiO2 in the g-C3N4 structure. In all samples, the sharp PL emission peak at 455 nm can be observed. The intensity of the PL peak reduced as the g-C3N4 content ratio decreased [122]. In addition, as can be seen in Figure 10f, the electron resistance decreased when g-C3N4-TiO2 heterojunction was used. Additionally, stability is another vitally important factor in determining the photocatalytic activities, and it is shown that the g-C3N4-TiO2 heterojunction demonstrates excellent stability after three to five cycles under different circumstances [117,120,122,123].
Many factors may affect the g-C3N4-TiO2 heterojunction productivity and improve the photocatalytic activity by enhancing the charge carrier separation and prevent recombination [124,125,126]. Rathi and coworkers showed that the CuNi@g-C3N4-TiO2 nanocatalyst had a 3-fold and 5-fold higher photocatalytic activity than bare g-C3N4 and TiO2 nanorod for Rhodamine B degradation. Besides, the photocurrent density of the TiO2 nanorod, bare g-C3N4, Cu@g-C3N4, Ni@g-C3N4, and TiO2/CuNi@g-C3N4 is 0.108 mA/cm2, 0.377 mA/cm2, 0.530 mA/cm2, 0.6012 mA/cm2, and 0.890 mA/cm2, respectively. It should also be mentioned that the charge separation was promoted since the presence of Cu and Ni species [126]. In another work, researchers also prepared Ti3+-TiO2/O-g-C3N4 heterojunctions via a hydrothermal approach [125]. In this synthesis approach, 1 g g-C3N4 should disperse with titanium oxohydrides sol precursor at room temperature for 20 min, ultrasonically. The collected sample should transfer into the Teflon-lined autoclave at 160 °C for 27 h and then be washed and dried at 60 °C for 3 h. This method was used to prevent the g-C3N4 aggregation and fabricate exfoliated g-C3N4 nanosheets. The synthesized heterojunction significantly decreased the regenerated electron-hole pairs’ recombination. Additionally, the conductivity is greatly enhanced and widens the light absorption range by adding the Ti3+, and O. P is another element used to improve the heterojunction connection and promote photocatalytic activity by facilitating the carriers’ transfer and separation [127,128].
Various novel metal nanoparticles are leading to the improvement of photoexcited semiconductors because the surface plasmonic effect is beneficial to reduce the photogenerated electron-hole recombination, improve the efficiency of visible light absorption and photocatalytic activity. Silver, and gold nanoparticles, for instance, have high stability and good conductivity. Au nanoparticles can increase the electron concentration onto their surface and enhance and extended adsorption for catalytic activity by its surface π bond [129,130,131]. Ag nanoparticles (AgNPs) are also used to modify the g-C3N4-TiO2 heterojunction [132,133,134,135]. The presence of AgNPs not only promotes the visible light response due to the surface plasmon resonance (SPR) effect, but also, they can result in capturing the electrons, to separate them, and transfer them more easily. These electrons react with the absorbed O2 on the AgNPs modified TiO2@g-C3N4 to form O2. With reference to the high specific area, electrical conductivity, and mobility, different graphene-based materials such as reduced graphene oxide [119,136]. Other materials can also be used to modified g-C3N4-TiO2 heterojunction to increase the specific area and separate the photo-induced electron-hole pairs to enhance photocatalytic efficiency [137,138,139].
The structure and morphological evaluation of the prepared g-C3N4–metal oxide-based heterojunctions have been investigated in different research works [140,141,142,143]. Jo et al. revealed structural properties of g-C3N3-TiO2 heterojunctions, which are consistent with other similar composites [144]. Figure 11a showed the transmission electron microscope (TEM) micrograph and the selected area electron diffraction (SAED) pattern of 5%-g-C3N4/TiO2 nanoparticles. Researchers observed that g-C3N4 nanolayer uniformly is covered with TiO2 nanoparticles, which suggested an intimate interface between them. Besides, the inset Figure 11a showed that the circular rings corresponded to the (101), (004), (200), and (105) planes of the polycrystalline TiO2 nanoparticles anatase phase. A TEM image showed that the TiO2 nanoparticles are deposited onto the layered structure on g-C3N4. Moreover, the high-resolution transmission electron microscope (HR-TEM) images of the 5%-g-C3N4/TiO2 nanotube indicates the interplanar distance of 0.350 nm, attributed to the (101) plane of anatase phase of TiO2 (Figure 11b,c). In addition, elemental mapping analysis of the 10%-CN/TNP (Figure 11d) suggested that TiO2 is uniformly present onto the g-C3N4 surface. Figure 11d also revealed the presence of Ti, O, C, and N, which suggested the co-existence of both layered g-C3N4 and TiO2 nanotube [144].

5.2.2. ZnO-g-C3N4

ZnO nanostructures such as nanosheets, nanoplates, and nanorods are semiconductors that have been used for preparing of g-C3N4-based heterojunctions. Like the TiO2, the ZnO bandgap is about ~3.2 eV with ECB and EVB of about 2.7 eV and −0.5 eV, respectively [145]. We showed that the ZnO nanostructure’s size, shape, and order could be tuned by an interplay of magnetic and gravity forces [146]. We also demonstrated the enhanced microbial detection capability when the synthesis was conducted under these external forces by changing the materials’ electrical resistance based on surface interactions. Because of low-cost of the preparation, having a large surface area, high aspect ratio, proper bandgap energy, minimal toxicity, and good stability, ZnO nanostructures have captured considerable researchers’ attention [147,148]. However, ZnO suffers from minimal light absorption (5% of the ultraviolet spectrum of the sun energy). Additionally, a high electron-hole recombination rate is another undesirable factor of ZnO for photocatalytic applications.
To deal with these problems the heterojunction of ZnO with g-C3N4 might be an option for different photocatalytic applications because, coupling g-C3N4 with ZnO nanostructures can improve charge migration, separation and prevent electron-hole recombination. Researchers have reported various methods for preparing the g-C3N4-ZnO structure, such as hydrothermal, solvothermal, atomic layer deposition, etc. [149,150,151,152,153]. Jung et al. synthesized g-C3N4-ZnO with various thermal treatment and condensation temperatures (T = 350, 400, 450, 500 °C). The BET specific area and ZnO crystallite size of the prepared Z-scheme g-C3N4-ZnO structure is decreased by increasing the thermal treatment temperature [154]. Other researchers who designed g-C3N4-ZnO via thermal treatment announced that the optimal amount of the g-C3N4 content in the composite is 5.0 wt.% [155]. Melamine and ZnCl2 should be vigorously stirred for 20 min in a 250 mL beaker, then, Na2CO3 is added dropwise into the suspension and stirred magnetically for 30 min, and finally, dried at 60 °C for 30 min. The product was placed into a crucible with a cover to prevent from the volatilization of melamine and heated at 500 °C for 2 h at a rate of 10 °C/min. The g-C3N4/ZnO photocatalysts were obtained after deamination treatment at 520 °C for 2 h. The effect of different g-C3N4 precursors such as dicyandiamide (DCDA), urea, and thiourea on the g-C3N4 and ZnO interaction and structural morphology was investigated in another work [156]. The excellent interaction of the DCDA-ZnO results in the perfect Z-scheme charge transfer core-shell structure with the ZnO core and shell of the g-C3N4, and the low electron density of the PL emission resulting in promoting the efficiency of the Methylene Blue (MB) photocatalytic degradation. If the interaction between the precursors, such as urea and thiol, and ZnO is weak, the porous, segregated morphology is obtained. Hydrothermal method to prepare g-C3N4-ZnO heterojunction [149] not only is a low-cost preparation but also Zhang et al. revealed that it could detect nine pesticide residues in four different samples simultaneously. The solvothermal synthesis method of the preparation of the g-C3N4-ZnO modified the TiO2 nanotube arrays by using the ethylene glycol solution is reported by Mohammadi et al. [150]. Zhang and coworkers synthesized the g-C3N4-ZnO heterojunction composite in the metal ion-containing ionic liquid’s presence by the solvothermal method [156]. g-C3N4 was added to the ZnCl4 in an ethanol solution, sonicated, and mixed with NaOH. Then, the mixture should be placed in a 25 mL Teflon-sealed autoclave and maintained at 160 °C for 24 h. After washing with distilled water and absolute ethanol, the dried g-C3N4-ZnO powder was provided. The strong interaction between g-C3N4 and ZnO result in the higher migration of the generated electrons and slower recombination rate was prepared. They revealed that the power conversion of the structure composed by the solvothermal approach compared to the pure TiO2 nanotube arrays increases from 1.04% to 2.45%. Besides, the uniform type II heterojunction between g-C3N4 and ZnO can also be synthesized by atomic layer deposition (ALD) [151]. The mentioned heterojunction composite, synthesized by the ALD method, possessed a stable dispersion of g-C3N4 powder in the reactor, which prevented the charge carrier recombination. Mechanochemistry (mechanical milling) is also employed to prepare g-C3N4-ZnO composite [157]. This method provided significant photocatalytic stability and enhanced the composite’s photocatalytic activity, which was 3-fold higher than the bulk g-C3N4 because of the strong interaction with ZnO. Figure 12 illustrates the schematic of the ZnO/g-C3N4 preparation method.
Introducing the ZnO to the g-C3N4 is proved that this structure can enhance photocatalytic performance such as charge transfer and separation and decrease the photogenerated carriers’ recombination (Figure 13a) [159,160,161,162,163]. In recent years, several types of research have been worked on the characterization of these heterojunctions. For example, Wang et al. demonstrated the eight XRD characteristic peaks for the pure Zn in the g-C3N4-ZnO sample, shown in Figure 13b [164]. The reduction in the g-C3N4 content results in a decrease in the intensity of its two prominent characteristic peaks. Besides, they also revealed the FTIR analysis of the synthesized composites. The g-C3N4-ZnO composites’ FTIR peaks show the main peaks of the bulk g-C3N4, which are shifted to the lower wavenumber, and this is because of the low strength of the characteristic bonds. The FTIR peaks of the g-C3N4-ZnO heterojunction are also similar to those of the main peaks of the g-C3N4 wavenumber. The indication announces a chemical bond in the heterojunction between g-C3N4 and ZnO so that this structure will improve the charge transfer and photocatalytic efficiency. Moreover, the UV-vis DRS spectra of the g-C3N4, ZnO, and g-C3N4-ZnO heterojunction with the different g-C3N4 ratio is illustrated in Figure 13c [164]. The absorption edge of the g-C3N4 is about 460 nm, attributed to the bandgap of 2.69 eV. Besides, the absorption wavelength of the pure ZnO appeared at 396 nm showing the bandgap of 3.13 eV. Finally, the g-C3N4-ZnO revealed the redshift for the higher g-C3N4 content, which is extended to the visible-light region. The intense visible light absorption is another significant indication of the strong chemical interaction and more electron-hole generation leading to the photocatalytic activity’s improvement [164]. The high photocatalytic activity of the mentioned structure is mainly used to degrade of dyes such as malachite green (MG). After the reaction of the photocatalyst with MG, the photocatalytic rate in 0, 5, 10, 15, 20, 25, 30, 35 and 45 min was measured (Figure 13d) [165]. A similar analysis was conducted on other dye types, such as Rhodamine-B (Rh–B), Congo red (Con-R), and Red ink (RI) solution. The results show the degradation efficiency after 45 min in the presence of the g-C3N4 was determined ~97.24%, 82.37%, 70.05% and 46.99% for MG, Rh-B, Con-R, and RI’s degradation, respectively [165]. Figure 13e demonstrated the photoinduced charge transfer capability of g-C3N4-ZnO nanorod arrays. The photocurrent density of the prepared materials was generated and increased under visible light irradiation. However, the photocurrent density decreased as the illumination was stopped. It was also revealed that the g-C3N4-ZnO produced the most photocurrent density compared to the bare ZnO and g-C3N4, showing excellent charge transfer and separation. The charge transfer resistance is depicted in Figure 13f, shown impedance spectroscopy (EIS) Nyquist plots [160]. The arc radius in EIS spectra corresponds to the resistance of the interface layer at the photocatalyst surface. As observed in Figure 13f, the smaller arc radius is shown for g-C3N4-ZnO than that of g-C3N4 and ZnO. To be more specific, the smaller arc radius leads to the lower charge transfer resistance, resulting in the improved photogenerated transformation at the interface. PL spectra announce the photocurrent recombination, separation, and migration rate [160]. The ZnO PL emission wavelengths under the excitation wavelength of 350 nm were low-intensity emissions at 445 nm and 600 nm with low intensity. Besides, the emission spectra of the g-C3N4 and g-C3N4-ZnO heterojunction under the same excitation wavelength are 452 nm and 506 nm, respectively. Moreover, the higher PL intensity of the composite than that of g-C3N4 and ZnO reveals a more recombination rate of photogenerated charge carriers via the Z-scheme pathway. One of the most critical factors in photocatalytic efficiency is the catalyst’s lifetime. Recyclability of g-C3N4-ZnO is obviously illustrated by the negligible loss of the photocatalytic activity after the five cycles resulting in good stability of the g-C3N4-ZnO photocatalyst [166,167,168].
The doping of the heterojunction is a method to change and tune the properties of the bare g-C3N4-ZnO heterojunction. The metal elements are one of the excellent candidates used as dopants in the g-C3N4-ZnO structure. Ahmad et al. illustrated the effect of different metal dopants on photocatalytic efficiencies [169]. Ag-doped g-C3N4-ZnO showed the highest specific area compared to the pure ZnO, g-C3N4, and Al, Mg, Ni, Cu-doped heterojunctions. The redshift is observed for UV-vis absorption for the metal-doped g-C3N4-ZnO composite compared to the pure ZnO, revealing the more visible light absorption, and inducing a higher charge carrier generation rate [169]. The bandgap of the ZnO, g-C3N4, Al, Mg, Ni, Cu, and Ag-doped g-C3N4-ZnO, which is determined by the Tauc’s plots, is 3.23 eV, 2.66 eV, 3.15 eV, 3.08 eV, 3.06 eV, 3.0 eV and 3.05 eV, respectively. The PL spectra of these doped composites demonstrated the three emissions at 391 nm, 456–466 nm, and 550 nm, which are attributed to the free exciton transition or recombination process, bandgap recombination of photoinduced charge carriers, and oxygen vacancies [169]. The photocurrent of the metal-doped structure is higher than that of the pure ZnO and g-C3N4. The Cu-doped composite showed the lower electron transition resistance among all composites. It was concluded that the synergistic impact of Cu-doped g-C3N4-ZnO, promotes electron mobility and separation efficiency [169]. Photocatalytic activity of the Mg-doped composite is higher than that of ZnO, g-C3N4, and g-C3N4-ZnO structure. Mg is one of the ideal candidates for charge separation in the composite structure [170]. Other kinds of metal-based materials such as K, Cr, Co, and Fe have shown excellent light absorption, increased photo-induced electron-hole generation and separation, and migration for improved photocatalytic activities [171,172,173,174]. ZnO/K@g-C3N4 had higher stability after five cycles (84%) for tetracycline removal. Doping the Cr into the g-C3N4-ZnO structure promoted photocatalytic performance [172]. 60% g-C3N4/Cr-ZnO photocatalyst had 93% degradation rate in 1.5 h, which is 3.5, 2.5, and 2-fold higher than that of 5% Cr-ZnO, bulk g-C3N4, and 60% g-C3N4-ZnO, respectively. Nitrogen is another widely important element used as a dopant for the g-C3N4-based heterojunctions [147,175,176,177]. g-C3N4 with 7 wt % of N-ZnO showed the highest photocatalytic RhB degradation with 5 and 4 times higher than those of N-ZnO, and g-C3N4, respectively [175]. g-C3N4-ZnO with N doping exhibited a 77% higher H2 evolution rate than that of pure carbon nitride and showed a great charge carrier transfer [147,176]. The N-doped heterojunction revealed the high photocatalytic activities for methylene blue degradation since it shows the narrower bandgap [178]. Other researchers showed the impact of carbon doping in the ZnO-g-C3N4 composites [177,179,180,181]. The Z-scheme heterojunction system containing C-doped g-C3N4 grafted on the C, N co-doped ZnO was used to improve the optical properties for enhancing the BPA organic pollutants photodegradation and hydrogen evolution reaction [177]. Oxygen and sulfur are also used to improve the efficiency of the composites [181,182].
Different g-C3N4-ZnO heterojunction-based ternary composites have been constructed by various researchers worldwide [183,184,185,186,187]. The surface plasmon resonance (SPR) effect of Ag and Au nanoparticles (NPs), resulting in the improved photocatalytic activities improve by increasing the electron-hole generation and separation, so many researchers tend to use Ag NPs in the heterojunction structures [188,189,190,191,192]. Ag NPs facilitate the migration and improve the photoinduced electron-hole pairs separation by creating close interfaces between g-C3N4 and ZnO. Besides, Ag NPs reduce the energy barrier for CO2 to increase the intermediate radicals on the surface of the nanocomposites [191,193]. It is also revealed that the reaction constant rate for Ag (5 mol%)/ZnO/g-C3N4 is 2.4 times higher than ZnO/g-C3N4. Ag (5 mol%)/ZnO/g-C3N4 composite extends its surface, leading to promoting the photogenerated electron-holes pairs and increasing the lifetime and stability of the charge carrier [194,195]. To show the improved photocatalytic performance with high stability, different Ag-based compounds between g-C3N4 and ZnO have been investigated [188,196,197,198,199]. In another research, the hydrothermally synthesized Ag-ZnO/S-g-C3N4, comprising ZnO NPs doped with 7% Ag with 25% Sulfurized-g-C3N4, exhibited outstanding MB photodegradation (97% in 40 min) with excellent recyclability [200]. Carbon-based materials are also used beside the g-C3N4-ZnO structure [199,201,202]. Graphene oxide (GO) is carbon-based material commonly used in the g-C3N4-ZnO heterojunction [199,203,204]. GO is a two-dimensional platform, providing a promising electron conductivity, high specific surface area, and Young’s modulus, which are useful for improving photocatalytic activities. The RhB dye degradation is about 99% for ZnO-g-C3N4-GO nanocomposites in 14 min [199]. The trinary nanocomposites provide high stability, which can be used for a wide range of environmental applications [203]. Other forms of carbon-based nanomaterials such as carbon dots (C Dots) are also used to improve the heterojunction structures’ efficiency [201,205]. The C Dots provide the facile photoinduced electrons transfer from the ZnO’s CB to the g-C3N4′s VB. The Z-scheme heterojunction structure can be used in biomedical applications for bacteria-killing and acceleration of wound healing system [201].

5.2.3. Iron Oxide-g-C3N4

FexOy such as FeO, Fe3O4, and Fe2O3, not only can improve the photocatalytic performance of the g-C3N4structure due to some unique characteristics but they can also be vastly used for contaminant removal in various media [206,207,208,209]. Compared to the combination of g-C3N4 with TiO2 or ZnO, fewer papers have been focused on the g-C3N4-FeOx heterojunctions. Xu et al. announced that the CB and VB of Fe2O3 are 0.3 eV and 2.4 eV, respectively [210]. FeOx failed to show any appreciable photocatalytic activity since the improper, more positive CB edge position (Figure 14a). It is confirmed that g-C3N4/α-Fe2O3 nanocomposites obey a Z-scheme mechanism for photogenerated charge separation. Working on the g-C3N4-FeOx composite was ignited by the pioneering study authored by Ye et al. [211]. They demonstrated that the efficiency of the Fe2O3/g-C3N4 photocatalysts was increased up to 1.8 times than the bulk C3N4 for the RhB degradation under visible light irradiation.
One of the most common synthetic methods to make iron oxides, especially superparamagnetic iron oxides (SPIONs), which show the highest saturation magnetization among all iron oxides while having a low magnetic coercivity, is co-precipitation method [212]. The effect of different parameters such as solution temperature, alkalinity, and stirring rate were investigated by Hosseini’s group on the suspension properties for MRI applications [213]. Iron oxides have many applications in different area such as biomedical applications, environmental application, etc. [214,215]. Wang et al. prepared g-C3N4-Fe2O3 nanocomposite via chemical co-precipitation approach [209]. They also evaluate the adsorption and desorption of seven polycyclic aromatic hydrocarbons (PAHs). Low limits of detection (LOD), excellent linearity, and recovery of the g-C3N4/Fe3O4 nanocomposites revealing their ideal candidacy for environmental applications, especially PAHs removal from water samples. In another research, the Fe2O3-g-C3N4 heterojunction was synthesized by thermal treatment in a hypoxia environment [216]. The presence of Fe2O3 in the structure will reduce the recombination rate and promote N2 adsorption. Some other researchers used in-situ thermal condensation to prepared g-C3N4 with iron oxides, which can be used in ciprofloxacin (CIP) degradation [217]. CIP is an antibiotic for bacteria sterilization by inhibiting the bacterial DNA. Besides, it is demonstrated that the remnant CIP in the soil can be absorbed by plants and transfer to the human body by consumption. The long-term CIP intake leads to some serious health issues [218]. The hydrothermal method is another method for preparing this composite [219]. In this method, researchers used the solution containing colloidal of the mixture of both Fe2O3 and g-C3N4. Fe2O3 could not cause the methanol yield since the low conduction band of the Fe2O3. To be more specific, Fe2O3 (5, 10, 15, 20 wt %) was added to 500 mg of g-C3N4 (in 25 mL of water), prepared by direct solid-state reaction of dicyandiamide and thiourea, to form a homogeneous mixture. The solution was transferred into the Teflon-lined stainless-steel autoclave at 150 °C for 4 h, and then washed and dried at 60 °C overnight. As a result, Duan and Mei revealed that the g-C3N4-Fe2O3 heterojunction significantly improved methanol yield from the CO2 photoreduction.
As mentioned above, Fe2O3 has a narrow bandgap of about 2.1 eV, making the iron oxides an excellent candidate for broad visible light absorption. Some factors, such as the high recombination rate of electron-holes, the short diffusion length of holes, and lack of sufficient conductivity, improve the need to constructing a FeOx-based composite, which will increase the applications of this composite. One of the most well-known FeOx-based composites is the g-C3N4-Fe2O3, which can adequately promote photocatalytic activities. Geng et al. illustrated the XRD and FTIR pattern of the g-C3N4-Fe2O3 composites, which are similar to the results of other publications [220]. Any significant peak shift is observed in the XRD pattern of the composites, which are indicated in Figure 14b. Additionally, the intensity of the peaks related to the (104) and (110) of α-Fe2O3 are strengthened by increasing the α-Fe2O3 content in the structure [220]. They also investigated that the FTIR characteristic peaks of the g-C3N4-Fe2O3 composites are not significantly change compared to the bare g-C3N4 and Fe2O3 (Figure 14c). As mentioned previously, the broad peak around 3500 cm−1 is due to the presence of moisture in the samples. The characteristic peaks at the bare Fe2O3 and g-C3N4 comprising 3000–3400 cm−1 and unresolved peaks from 1237 to 1640 cm−1 indicating N–H, C–N, and C=N in the g-C3N4, respectively [220]. Besides, the sharp peak at 811 cm−1 is related to the stretching vibration of the triazine units. At the same time, the peaks at 543 cm−1 and 469 cm−1 directly correspond to the Fe-O stretching vibrations. The optical studies are also depicted in Figure 14d,e, which not only showed absorption and emission of this composite but also revealed the effect composition on these characteristics features. All samples showed a promising UV absorption above 450 nm compared to the g-C3N4, helping the photocatalytic characteristics improvements (Figure 14d) [221]. The PL emission spectra showing in Figure 14e will aid us in a clearer understanding of the photophysical behavior of the prepared materials. The results showed a broad peak at the range of 440–463 nm, while the PL intensity of the FexOy showed a high decrease. The decrease in the PL intensity may be due to the reduction in the luminous recombination probability, resulting in enhancing the charge separation and photocatalytic reactions. Cheng et al. also demonstrated the insignificant deactivation in about 16 h during 4 cycles showing high stability of this composition. In order to improve the photocatalytic activities, Wang and his coworkers used Al–O bridged g-C3N4-α-Fe2O3 z-scheme nanocomposites [222]. In this work, the photocurrent density with 450 nm excitation wavelength increased when researchers used the Al–O bridged 15 g-C3N4-α-Fe2O3, containing 15 mass percentage of g-C3N4 and 6 mole percent of Al to Fe. This analysis indicated in Figure 14f showed that Al–O bridged g-C3N4 facilitated the photo-generated charge transfer and separation [222].
Metastable materials that transform from one to another state over a long period of time have superior properties. Bibyite phase of iron(III) oxide (β-Fe2O3) act as its α-phase (hematite) iron(III) oxide. However, it shows a more desirable bandgap (1.8 eV) for photocatalysis. Christoforidis et al. prepared metastable β-phase Fe2O3 nanoparticles on the g-C3N4 surface by a solid-state, in-situ growth method, without the need of specialized equipment, surfactants, stabilization, or precipitating agents [223]. In this research, the β-Fe2O3 improve the photocatalytic activities by increasing the ability of light absorption in the visible region, and enhanced carriers’ separation. The hybrid β-Fe2O3/g-C3N4 nanomaterials are an excellent candidate for photodegradation since they showed higher photocatalytic activity and a promising stability [223].
There are several g-C3N4-FeOx-based composites with promising photocatalytic activities. The g-C3N4 coated with the FexOy covered by metals will enhance the photocatalytic properties by ameliorating the photo-induced charge generation and separation. The promising LSPR effect of the Au and active sites of Pt convinced researchers to use Fe2O3/Pt/Au nanocomposite immobilized on the g-C3N4 surface as an excellent composite for hydrogen evolution [224]. As discussed in our previous papers, Au nanoparticles with strong LSPR can participate in catalytic reactions (for an example see [225]). Au NPs with a strong LSPR effect enlarge the photo-electron conversion efficiency of the photocatalyst. The researcher also investigates the CO2 photoreduction of g-C3N4 quantum dots-Au NPs co-modified CeO2/Fe3O4 micro-flowers (MFs). They showed that Au NPs promote photocatalytic activities by generating the electrons and holes and enhance carriers’ separation in CeO2 MFs and CN QDs in the photocatalyst [226]. Copper is another metal used in the g-C3N4-FeOx-based composite. Liu et al. fabricated g-C3N4/Fe2O3-Cu for electrochemical detection of glucose [227]. g-C3N4/Fe2O3-Cu composites improved the electrochemical performance for glucose detection with a LOD of 0.3 mM. They also confirmed that g-C3N4-FexOy composite can be used as an electrode in sensors to measure other compounds. The effect of a multi-walled carbon nanotube was investigated by Zhang et al. [228]. They demonstrated that due to the large specific area, hydrogen bonds, π-π, and electrostatic interactions of the MWNTs@g-C3N4@Fe2O3, these 3D structures were novel magnetic solid-phase extraction sorbents for PAH with the LOD of 0.001–0.5 mgr·L−1. Besides, the 3D structure has a good repeatedly and recovery for 16 PAHs in the water samples. Graphene, as other carbon-based materials also used to improve the efficiency of Fe3O4/g-C3N4 composites. Wng et al. showed that the Fe3O4/graphene/S doped g-C3N4 dose of 1.0 g/L comprising 20% Fe3O4 mass fraction could completely remove Ranitidine (≤2 mg/L) in 60 min, an initial pH of 7.0 [229].
The other ternary composite used for the water splitting is g-C3N4/CeO2/Fe3O4 [230]. The composites showed enhanced oxygen and hydrogen evolution reaction with high current density (40 mA·cm−2) at the potential of 327 mV, which was greater than the bare Fe3O4 and bulk g-C3N4. The ternary composites showed excellent stability and negligible activity loss up to 14 h. the Z-scheme g-C3N4/Fe3O4 can be coupled with the CdS and was used for different antibiotics degradation [231]. The results showed the improved degradation rate to 45 times of CdS, 26 times of pure g-C3N4, and 9.5 times of CdS/g-C3N4 for the tetracycline removal. The presence of the Fe3O4 improves the photocatalytic performance and stability by an increase in the inducing and separation of the electron-hole pairs and generating more .O2 for organic pollutant degradation [231].
Other FexOy compounds are also used to improve the UV-visible light absorption for different applications. NaFe2O4 is one of the most well-known FeOx structures, using in the g-C3N4-based composites [232,233,234]. The Fe3O4@NiFe2O4-g-C3N4 improves photocatalytic activity up to 90% of CIP degradation by reducing the electron-hole recombination [232]. In all superparamagnetic samples, the high Ms assists in improvig of the recyclability and stability of the photocatalyst. The more magnetite content in the sample leads to a higher probability in the particle agglomeration, resulting in the decrease in the active sites of the and the reduction in the photocatalytic efficiencies. The fabricated g-C3N4/NiFe2O4 can also be used to degrade MB and RhB by activating H2O2 to produce the oxidizing reagent [233,234]. The magnetic properties of the NiFe2O4 (Ms= 45 emu/g) were induced to the whole composite (Ms= 40 emu/g) to promote repeatability and improve the photodegradation rate [233]. The other most popular compound captured researchers’ attentions are ZnFe2O4 and LaFeO3 [235,236,237].

5.2.4. WO3-g-C3N4

Figure 15a demonstrated the band edge position of the WO3 compared to the g-C3N4. This Figure illustrated that the bandgap of the WO3 is 2.6 eV with VB and CB of 2.9 eV and 0.3 eV, respectively, which are more positive than that of g-C3N4 and the O2/O2 potential [238,239,240]. Thus, it is approximately impossible to produce .O2 by using the traditional type-II mechanism. Many researchers showed that the charge migration almost occurred by the Z-scheme mechanism in the binary g-C3N4-WO3 composite [241,242,243]. Various methods have been fabricated this binary composite. The hydrothermal approach is one of the main routes to prepare the g-C3N4-WO3 composite [244,245]. Zhang et al. prepared WO3/g-C3N4 by dissolving the WCl6 and ascorbic acid in ethanol and g-C3N4 followed by 5 min sonication and stirred for 20 min [244]. The uniform suspension was heat-treated at 220 °C for 12 h, and then washed several times using ethanol. The other way for the preparation of the heterojunction is the microwave irradiation technique with the direct calcining of the WO3 and g-C3N4 combination at 400 °C for 2 h [246]. In this method, to prepare WO3 by simple household microwave irradiation, tungstic acid in NaOH solution was mixed and stirred for 30 min to form a tungstic hydroxyl group [246]. Afterward, the pH of the prepared mixture was reduced to 1 by adding HCl solution, and the gel should be put into the Teflon-lined household microwave oven (2.45 GHz) for 10 min to prepared WO3. Finally, the mixture should be placed in an alumina crucible with a cover in a muffle furnace and heated at 400 °C for 4 h. A wet chemical process, sonochemical, in situ self-assembly, etc., are other ways to synthesize the g-C3N4- WO3 heterojunction [247,248,249].
g-C3N4-WO3 composites were characterized by various techniques. In most research papers, XRD and FTIR are the prime characteristic methods to evaluate the formation of g-C3N4-WO3. Researchers announced that the WO3/CN wt % = 10% provides the best characteristics compare to the balk g-C3N4 and WO3 [250]. Figure 15b shows the XRD patterns of the bulk g-C3N4, WO3, and g-C3N4-WO3, perfectly showing the impact of WO3 content on this characterization. There are nine distinct peaks for the as-prepared WO3 [247,251]. Additionally, the binary composites revealed the combination peaks of the WO3 and g-C3N4. The higher WO3 content leads to the lower peak’s intensity of the g-C3N4, corresponding to the expansion of the interlayers and g-C3N4 coverage with a WO3. The low intensity of the peaks at 23.5° and 36.6°, which is detectible in the sample with a high WO3 to g-C3N4 ratio, is related to the hexagonal-phase WO3 and illustrated the formation of the composite successfully. They also characterized the g-C3N4-WO3 composite with the FTIR spectra analysis [251]. Like other materials’ FTIR spectra, the wide peak at the range of 3000–3500 cm−1 in all samples is attributed to the N-H and O-H stretching vibration, and the adsorbed water molecules’ bending vibration stands at about 1630 cm−1 for samples (Figure 15c). The broad peak at the 750–1000 cm−1 corresponds to the O-W-O stretching vibrations of WO3 [252,253]. As a result, although the g-C3N4 characteristic FTIR peaks can be observed in the composite FTIR spectra, WO3 peaks are not significantly detected in the hybrid composite, which can be ascribed to the vacancies between the g-C3N4 clusters or band overlapping. Optical studies were carried out by Chai et al. on the g-C3N4 with WO3 structure [254]. The UV-vis DRS of the WO3, g-C3N4, and WO3-g-C3N4 composites with the WO3 contents is depicted in Figure 15d. The obvious absorption edges at ~470 nm and 455 nm are detected for WO3, g-C3N4, respectively, corresponding to the 2.64 eV and 2.73 eV. The g-C3N4-WO3 composites present the combination absorption features of g-C3N4 and WO3 [254]. Figure 15e displays the PL emission peak of the bulk g-C3N4 and 18.6 wt % WO3/g-C3N4 hybrid composite [254]. The result demonstrates that the intensity of the composites’ PL peak at ~440 nm is lower than that of the pure g-C3N4 [255,256,257]. The PL spectra revealed that WO3 apparently suppresses the photoinduced electron-hole recombination in the WO3/g-C3N4 composites and confirms the Z-scheme interface contact [258]. They are also revealed that the OH generating during the photocatalytic reaction leads to the higher PL intensity increases the irradiation time. Besides, the transient PL decay trace of the bare g-C3N4, WO3, and g-C3N4-WO3 composite hollow microsphere are ~ 4.46 ns, 1.62 ns, 2.23 ns, respectively [259]. The zeta potential values of the WO3, g-C3N4 and their binary composites announce the negatively charged surface of the samples [260]. The presence of WO3 in the composite changes the zeta potential from −5.7 mV to −33.1 mV, while it decreases the BET specific area from 100.97 m2·g−1 to 47.88 m2·g−1, which are high enough for promising adsorption ability and photocatalytic activities. g-C3N4-WO3 reveals high stability and repeatability after 4 cycles with a slight efficiency decrease for the degradation of RhB [261,262,263].
There are also some dopants and compounds used to improve the photocatalytic efficiency of the binary composites. In the z-scheme C or Pt-g-C3N4 with hydrogen treated WO3, the electrons from g-C3N4 and holes from WO3 facilitate the photogenerated charge carries generation, which will enhance the photocatalytic activities [254,264]. In other words, the C or Pt dopants help the composite to improve the light absorption and charge separation. This structure also enhances stability and repeatability. O-doped g-C3N4-WO3 has 4.7 times higher H2 performance than the bare composite [265]. The carbon vacancies and g-C3N4 oxidization resulted in the formation of the porous composites and a decrease in the composite’s surface area. The effective Si-O bridge between g-C3N4 and WO3 significantly promotes charge transfer and separation [266].
In addition to the doped binary composites, adding other compounds to form the ternary structure is a practical way to boost the composite characteristics. Mediators and co-catalyst such as metals (such as Ag, Au, Cu, Pt, and Sn) are currently used to improve the Z-scheme composites by facilitating the transporting and carrier capturing during the photocatalyst process [267,268,269,270,271,272]. Li et al. demonstrated the charge carrier migration between various WO3-Metal (Cu, Ag, Au)-g-C3N4. They showed that Cu plays the ideal candidate for photocurrent enhancement and improves the photocatalytic performance in the Z-scheme g-C3N4-WO3 heterojunction among different metals used in the research. In other words, the Cu-g-C3N4 and WO3-Cu are favorable for electron migration since they have the matched Fermi level of energy [267]. In other research, the developed WO3/Ag/g-C3N4 ternary composite was used for the RhB and tetracycline (TC) degradation. The results demonstrated that the improved photocatalytic activity of WO3/Ag/g-C3N4 is obtained due to the large contact region between g-C3N4 nanosheets and WO3 nanoplates. Besides, the presence of the Ag NPs in the composite with SPR effect accelerates the charges to transfer, improves the photocatalytic activity, and enhances the stability and repeatability [268,269].
Researchers also worked on the Ag/g-C3N4/WO3 to degrade oxytetracycline hydrochloride under visible light [273]. The 0.4 g/L of Ag/g-C3N4/WO3 composite demonstrated the highest photocatalytic activity, which could degrade 97.74% of oxytetracycline (10 mg/L) in 60 min. In another research, Qin et al. showed that the presence of Pt in the WO3-g-C3N4 composites possesses excellent photocatalytic H2 evolution with 1299.4 μmol under visible light, which is higher than that of WO3/g-C3N4/Pt and pure CN with 1119.4 μmol, and 113.2 μmol, respectively [274]. As mentioned above, g-C3N4/Au/WO3 as Z-scheme heterojunction displayed excellent photocurrent that can be used in photoelectrochemical immunosensing of Aflatoxin B1 in food, which can be dangerous for humans [275]. The methane formation over the WO3-Bt-g-C3N4 composite was 5.98, 6.74, and 25.19 times higher than that of WO3-g-C3N4, Bt-g-C3N4, and g-C3N4 samples, respectively [272].
Some promising properties of tungsten-based materials such as electrical, optical, magnetic, and photocatalytic activities, low-cost preparation persuade researchers to use these materials such as NiWO4, BaWO4, CuWO4, and BiWO6 [270,276,277,278]. The band structures of these compounds with g-C3N4 and WO3 make the CB and VB positions match each other, leading to the prolonged charge carriers’ lifetime in the form of a double Z-scheme system. The photoelectrochemical (PEC) of a double Z-scheme g-C3N4-WO3-Bi2WO6 system reveals the enhanced photocurrent density, reduced electron-hole recombination, resulting in the promotion of photocatalytic efficiency [276]. Other materials have been anchored in the g-C3N4-WO3 to form an excellent composite for different applications [279,280]. Bi-based materials are among the prime and important materials used to construct the composites for photocatalytic activities [281,282,283]. MoS2 and MoO3 are other compounds to be utilized as materials for this ternary composite [284,285]. g-C3N4/WO3 with TiO2 can also be useful for the methylene blue dye degradation, which its efficiency is about 3.1 folds higher than that of the binary counterparts of TiO2/WO3 (0.00691 min−1) in 120 min. The kinetic constant for different composition of this composite are in order of TWG-15% (which 15% is the g-C3N4 wt %) > TWG-10% > TWG-5% > TWG-20% > TW [286].

5.2.5. Tin Oxide-g-C3N4

Other metal oxides widely used in g-C3N4 binary heterojunction are tin oxides. SnO (stannous oxide) and SnO2 (stannic oxide) are two forms of tin oxide. Figure 16a shows the band edges’ position in the tin oxide and compares it to the CB and VB of the g-C3N4. He et al. showed that the SnO2 has a large bandgap of about 3.6 eV, and its CB level is lower than that of g-C3N4 [287]. It is also demonstrated that these tin oxide-g-C3N4 composites mostly form the Z-scheme heterojunction [288,289]. Tin oxide can be designed with various shapes by wet and dry methods. The heterojunction is synthesized by different synthetic methods, including hydrothermal and thermal treatment [78,290,291,292]. Sadrnezhad and coworkers prepared g-C3N4-SnO2 with the pyrolysis of urea under microwave irradiation [293]. In this work, tin, ammonium, and urea were put into the beaker and placed in a microwave oven operating at 2.45 GHz and 900 W for 30 min. Finally, the product was washed and dried in the oven. In another work, the mesoporous SnO2 decorated with g-C3N4 was prepared via pulsed electrophoresis and facile water-crystallization [293]. With pulsing electrophoresis optimized parameter, the 0D g-C3N4 can be homogeneously and completely distributed inside the 1D SnO2. Sol-gel is also used to fabricate an efficient g-C3N4-SnO2 photocatalyst [294].
Several techniques have been used to confirm the formation of this composite. The XRD pattern and FTIR spectra of g-C3N4, SnO2, and SnO2-g-C3N4, are depicted in Figure 16b,c [295]. SnO2 does not change during the preparation of the final structure, which can be seen from the insignificant differences between the pure SnO2 and the SnO2-g-C3N4 XRD pattern in Figure 16b. The characteristic peaks in SnO2 revealed the (110), (101), (200), and (211) planes of a tetragonal rutile-like structure. Other small peaks, which are not mentioned in Figure 16b, are related to the (220), (310), (301), (202), and (321). As shown in Figure 16b, the (002) crystal plane peak of g-C3N4 overlapped with the (110) crystal plane peak of SnO2. Additionally, the Sn–O–Sn anti-symmetric stretching vibration between 400 cm−1 and 700 cm−1 was demonstrated in Reference [295] (Figure 16c). The red-shift peak at 597 cm−1 was a good indication of the formation of SnO2-g-C3N4 heterojunction [296,297]. The presence of SnO2 in the structure suppresses the electron-hole recombination rate. Peng et al. demonstrated the negligible differences and slight blue shift of g-C3N4-muscovite sheet/SnO2 structure, compared to the absorption edges of g-C3N4 (Figure 16d) [298]. Introducing the muscovite sheets does not affect the composites’ absorption, and g-C3N4-muscovite sheet/SnO2 absorption spectra reveal the potential photocatalytic applications under visible light. They also showed that the g-C3N4-muscovite sheet/SnO2 Cement has an excellent photocatalytic activity for RhB stains and isopropyl alcohol photodegradation. Due to the low electron-hole recombination probability of the SnO2-g-C3N4, the intensive PL intensity was observed in Figure 16e [298,299]. The faster interfacial charge-transfer and lower resistance, results in the Nyquist plot diameter decrease compared to SnO2 and g-C3N4, shown in Figure 16f [300]. This is therefore a promising photocatalyst. The properties and morphology of this composite change when the relative content of g-C3N4 varies in the heterojunction [301,302]. Chen et al. showed that the g-C3N4 mass ratio of 72% provides the highest photocatalytic performance, which is 17 times higher than bulk g-C3N4 [303]. SnO2-g-C3N4 heterojunction reveals notable stability and recyclability [18]. In addition to robust photocatalytic activities, this composite has a noticeable photoelectrochemical performance under visible light. Doping different elements, such as Sb, S, B, and P, is one of the main approaches to improve the performance of the SnO2-g-C3N4, which can modify the band edges [304,305,306].
Doping is a method to improve the efficiency of SnOx-g-C3N4 heterojunctions, which are used in different applications. SnO2/g-C3N4 sensor doped with Ni shows LOD about 1.38 ppb and has high stability, promising selectivity, rapid response and recovery time, and great resistivity against humidity [307]. Besides, researchers used made Ce doped SnO2/g-C3N4 to make capacitors with a specific capacity of <274 F/g. A supercapacitor with energy and power densities of 39.3 W h kg−1 and 7425 W kg−1, respectively, was made by using Ce-SnO2/g-C3N4/Activated Carbon. The supercapacitor device exhibited retention of 84.2% after completing 5000 cycles [308].
Additionally, several researchers used the SnO2-g-C3N4 based ternary composites to increase the applications of these kinds of composites [309,310]. The plasmonic Au-SnO2-g-C3N4 photocatalyst was fabricated for H2 evolution and degradation of the organic pollutant, which is commonly used due to the outstanding photocatalytic activities and significant stability (remain unchanged after 5 h in 5 cycles) [309,311]. In this ternary heterojunction, the presence of the plasmonic Au and g-C3N4 offer enhanced photogenerated electrons to the structure. SiO2 is another compound used besides the SnO2-g-C3N4 system for pollution treatment applications [312]. As mentioned above, TiO2 has promising properties that aid us in promoting the properties of this composite [313,314]. This ternary composite also demonstrated the perfect antibacterial activity for the degradation of E. coli bacteria, probably due to the interface between g-C3N4-SnO2 and TiO2 and lower charge recombination rate [314]. SnO2/chitosan/g-C3N4 nanocomposite has been used as an Electrochemiluminescence aptasensor to improve lincomycin detection [315]. Ali et al. also showed the cost-effective prepared g-C3N4/rGO/SnO2 nanocomposite for RhB degradation. The optimal amount of this nanocomposite reveals an increased RhB degradation efficiency [316].
As a result, the SnO2-g-C3N4 structure can be used in a wide range of applications. The exploitation of new clean energy instead of fossil fuels is of great interest to many since the fossil energy is exhausted. As a case in point, water splitting is a perfect example of this clean energy generation, which is widely applied by the g-C3N4-SnO2 [317,318]. Besides, in order to reduce environmental pollution, the demand for electric transports has been increased. Thus, it is essential to use the rechargeable batteries such as lithium-ion batteries (LIBs) and improve lithium storage capacity. Specifically, g-C3N4 enables SnO2 anode to enhance the Li storage in these batteries [319,320]. This heterojunction is also used to detect different compounds. Cao et al. showed that SnO2/g-C3N4 composite promoted the sensitivity and selectivity in ethanol gas-sensing applications [321]. The degradation of the inorganic pollutant is another application of this heterojunction [322,323]. The emission of nitrogen dioxide (NO2) and nitric oxide (NO) causing some environmental issues is one of the biggest challenges among several researchers, solved by Zou and coworkers by using SnO2-g-C3N4 photocatalysts using visible-light irradiation under 30 min [287]. SnO2-g-C3N4 is also used to decomposed Ammonium Perchlorate (AP), a toxic inorganic material [324].

5.2.6. Other Metal Oxides

Other kinds of metal oxides such as V2O5, NiO, MoO3, Cu2O, Co3O4, CeO2, Bi2O3, Al2O3, etc., are also used to improve the performance of the bulk g-C3N4 [325,326,327,328,329,330,331,332,333,334,335,336,337,338,339,340,341,342,343,344,345]. The g-C3N4-based heterojunctions can be modified by combining with several metals or doping with various agents [346,347,348]. Cu is a conventional metal using for the improvement of photocatalytic performance. Zhou et al. used Cu/Al2O3/g-C3N4 for Rhodamine B degradation by H2O2 [349]. The Cu immobilized Al2O3/g-C3N4 also showed promising stability for the treatment of water pollution. Besides, copper is used as a charge separation center for hydrogen evolution, MO, and phenol solution degradation under visible light [350,351]. In addition to Cu, the noble metal Ag and Au is another metal catalyst used besides the g-C3N4-based composite [347,352,353,354,355,356]. The Ag and Au can prevent rapid recombination probability, improve the transfer of a generated electron, and enhance the visible light absorption by the surface plasmon resonance, and can be used for a wide range of applications, especially decontamination of organic pollutants. Bi, Pd, Pt, Ni, and Cd are other metal photocatalysts used for improved photocatalytic activities [357,358,359,360,361,362,363]. Some semiconductors are also used besides the metal oxide-g-C3N4 based composites. Carbon-based nanomaterials provide excellent stability, cost-effective synthesis, enhanced photogenerated electron reservoirs used in bioimaging and sensing, photocatalysis, electrocatalysis [214,336,364,365,366,367]. Xie et al. fabricated the carbon quantum dots modified with MoO3-g-C3N4 and demonstrated that this structure showed outstanding visible-light absorption used to degrade tetracycline (TC) from the environment [364]. Besides, among all carbon-based nanomaterials, reduced graphene oxide (rGO) has captured intensive attention [368,369]. Gong and coworkers illustrated that the charges transfer between g-C3N4 and Bi2Fe4O9 (BFO) was improved by using rGO. The presence of the rGO causes the separation of electrons and holes in the CB of g-C3N4 and the VB of BFO, respectively [354].
In addition to the mentioned metals and semiconductors, metal-organic frameworks (MOFs) are a novel class of porous and crystalline materials with a large surface area-to-volume ratio, high porosity, and tunable pore size that might improve the biosensor sensitivity. Since these structures are made from metal ions, clusters, and organic ligands, these materials can promote the separation and transfer of photoinduced electrons, making them a promising candidate for photocatalytic activities, such as organic pollutants degradation, water splitting, and CO2 reduction [370]. Cui et al. utilized the Fe-based MOFs, MIL-53(Fe), with Bi2O3 and g-C3N4 for the degradation of amino black 10B since this structure can enhance the visible light absorption range [371]. Like other heterojunctions, doping is another element to improve the performance of these composites and widen their applications [372,373,374,375,376].
Some g-C3N4-based ternary composites comprising the metal oxide compounds can promote photocatalytic activities [103,377,378,379,380]. Bismuth complex oxides are among the most efficient catalysts with layered structures beside the g-C3N4-based composites. Among various Bi-based compounds, Bi2O3 based catalysts have drawn significant attention in the different areas [381,382]. It illustrated that CuO2/Bi2O3/g-C3N4 nanocomposite reveals improved photocatalytic activities for decomposing of 2,4-dichlorophenol under visible light [381]. In another research work, Vattikuti et al. prepared Bi2O3/V2O5 photocatalysts anchored on the g-C3N4 nanostructure, which can be used for the phenol red (PR) pollutant degradation [383]. They also demonstrated that the efficiency of hybrid composites for the PR removal under the simulation solar light irradiation was higher than that of fabricated materials. Bi2O3/g-C3N4 heterojunctions were also conjugated with the BiPO4 [15]. g-C3N4/Bi2O3/BiPO4 hybrid exhibited the perfect photoelectric performance, which is mainly due to the high separation and photogeneration charges and can increment the oxidation/reduction rate. Bi2O2CO3 is another Bi contained material that shows promising photocatalytic activities [384,385]. Kumar and coworkers suggested novel magnetic g-C3N4/Bi2O2CO3/CoFe2O4 heterojunction with high visible light absorption for reduction of 4-nitrophenol into 4-aminophenol [386]. TiO2-g-C3N4 based ternary composites are widely used for photodegradation. Min et al. fabricated the Cu2O-TiO2/g-C3N4 hybrid composite for the organic dyes’ discolorations. Besides, this nanocomposite can also illustrate good performance for discolorations of RhB, MB, and MO within 3, 10, and 15 min, respectively [387]. TiO2-g-C3N4 composites are also anchored with CeO2 and metal to form g-C3N4-Men+1/CeO2-TiO2 for photooxidation of toluene [388].

6. Application of Metal Oxide-Based g-C3N4 Nanocomposites

6.1. Photocatalysts

6.1.1. H2 Generation via Water Splitting

These days, the demand for safe, efficient, and renewable energy resources instead of limited fossil fuel sources has increased among more and more people [14,184,389,390,391]. This replacement is an efficient remedy for global warming and greenhouse gases emission. The hydrogen energy content is in the range of 120 to 142 MJ kg−1, which is higher than that of hydrocarbon fuels. Thus, it is estimated that hydrogen will be responsible for 90% of energy production by 2080. As a result, H2 generation is a novel and environmentally-friendly research topic among many researchers [8,13,392,393,394,395]. One of the most recent hydrogen production techniques is the photocatalytic water splitting method via metal oxide-g-C3N4 heterojunctions by using prolific light sources [365,390,396,397,398,399,400]. For water splitting, the band position of the photocatalysts should be modified to provide the CB position more negative than the H2O reduction potential (0 eV vs. Normal Hydrogen Electrode (NHE)) for H2 generation and more positive than the H2O oxidation potential (1.23 eV vs. NHE) for O2 generation. To be more specific, the generated electrons are used for the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) via Equations (1) and (2), respectively, and the final water-splitting reaction is shown in Equation (3) [111].
Hydrogen evolution reaction (HER): 2H+ + 2e → H2, E° = 0.00 eV vs. NHE
Oxygen evolution reaction (OER): H2O → 1/2O2 + 2H+ + 2e, E°= −1.23 eV vs NHE
Overall water splitting: H2O → H2 + 1/2O2, ΔG° = −1.23 kJ mol–1
where NHE is the normal hydrogen electrode.
Generally, there are three steps for each photocatalytic reaction; initially, the semiconductor absorbs light with energy equal to or higher than the bandgap to generate the electrons and holes in the valance and conductive band, respectively. Then, the photoinduced electrons and holes are moved to the surface of the semiconductor to start the reaction. Finally, the charge carriers participate in the reduction and oxidation reactions on the surface of the photocatalysts.
Considering the band edges position of some of the metal oxide-g-C3N4 composites, which were noted previously, some heterojunctions are more anodic than that of H2O reduction potential to show excellent performance under visible light irradiation [111,401]. As a case in point, TiO2-g-C3N4 heterojunctions are widely used as an excellent photocatalytic for H2 evolution. Yan et al. demonstrated that the efficiency of the visible-light-induced H2 evolution of the binary composite comprising anatase TiO2 and g-C3N4 was enhanced, and this is due to the desirable photoinduced carriers’ separation [402]. In order to raise the heterojunction efficiency, other kinds of materials with various effective characteristics are also used. These materials (metals) or cocatalysts such as Ag, Au, Pt, etc., can host active sites for H+ reduction [403,404]. The loading Au and Ag would be significantly beneficial for this application because of their plasmonic characteristic. Marchal et al. illustrated that the optimized components ratios and contact quality in Au/(TiO2–g-C3N4) lead to the enhanced visible light absorption with the proper band positions for photogenerated charge carriers [403]. Besides, the presence of Au and Ag will promote the water splitting for H2 production by the improved charge separation rate. The H2 generation rate under sunlight irradiation as a function of the relative ratio of methanol as a sacrificial agent and TiO2/g-C3N4 is depicted in Figure 17a. It was mentioned that no H2 generation was observed for the Au-free g-C3N4 and TiO2. Furthermore, the best photocatalytic H2 production was observed for the 0.5 wt % Au/(TiO2–g-C3N4) (95/5) structure, using 1 vol% of CH3OH as a sacrificial agent (Figure 17a). Carbon quantum dots (CQDs) can also improve photocatalytic performance to effectively decompose H2O2 to H2O and O2. CQDs possess up-conversion fluorescence spectra, which could convert visible light to ultraviolet or near-ultraviolet light, resulting in excellent photocatalysis. Consequently, it is proven that the combination of C3N4, TiO2, CQDs is a great candidate for water splitting [405]. Other metal oxide-g-C3N4 based composites such as WO3-g-C3N4 and ZnO-g-C3N4 are used for H2 generation. Mahala et al. demonstrated that the prepared ZnO nanosheets decorated with g-C3N4 quantum dots composites on the fluorine-doped tin oxide (FTO) coated glass slide could be utilized as a photoanode for water splitting via PEC (Figure 17b) [406]. In another work, the effect of boron addition and carbon nitride content did increase the H2 evolution up to 85% compared to the bare TiO2, which is mainly due to charge carriers’ generation and separation [109]. It was also shown that the photoconversion efficiency of the low charge-transfer resistance of ZnO decorated with g-C3N4 is 2.3 times higher than that of pure ZnO [406]. As discussed previously, this composite had a high specific surface area, promising electronic conductivity, and excellent charge transfer interfaces and would be excellent candidates for the water splitting and H2 evolution [407]. Other materials with suitable properties can also be used to improve the H2 generation of g-C3N4–metal oxide-based composites [141,283].
In a plasmonic photocatalyst, a metal nanostructure with a size less than the wavelengths of the light, is embedded onto dielectric or semiconductor materials [408]. In such systems, the light can lead to a local electromagnetic field by generating localized surface plasmon resonance (LSPR) and hot carriers. The hot carriers move to conductive and valance bands and this phenomenon is called the LSPR sensitization [409,410]. The hot carriers can be useful for direct oxidation or reduction of chemical species. Some of the metals that showed plasmonic effects are gold (Au), silver (Ag), copper (Co), and platinum. Zhao et al. state that the LSPR of Au can enhance the light absorption and increase the number of photogenerated carriers in the Au/g-C3N4/CeO2 plasmonic heterojunction. The heterojunction was employed for Cr6+ reduction and oxytetracycline hydrochloride (OTH) catalytic degradation [411].
Metal oxide-g-C3N4 based ternary composites capture lots of attention among researchers [412]. For example, g-C3N4 (CN)/TiO2 (TO)/PbTiO3 (PTO) films were prepared by the sol-gel followed by the CVD method and were investigated for PEC water splitting [413]. Wang et al. showed the decreased resistance of the interface (RCT) of CN0.10/TO0.4/PTO compared to the pristine PTO and CN0.10/PTO, which facilitated the charge transfer and reduced electron-hole recombination [413]. The incident photo-to-current conversion efficiency (IPCE) was calculated to measure the PEC performance. The pristine PTO has a lower IPCE than CN0.10/PTO, which is mainly due to the higher charge recombination rate. After the TO buffer layer’s insertion, the IPCE value was drastically increased to 14.2% at 380 nm [413]. The inserted compact TiO2 buffer layer provided the type II and Z-scheme interfaces between PTO, TO, and CN and promoted the PEC performance with an improved current density of −68.5 μA cm−2 at 0 V versus Ag/Ag-Cl electrode. The high performance is mainly due to the high photogenerated ability of g-C3N4/TiO2 heterojunction. BiVO4 is another cost-effective compound with a proper bandgap (2.4 eV) to enhance the PEC water splitting [184]. Like PbTiO3, BiVO4 can also improve the photocurrent density of the g-C3N4@ZnO/BiVO4 heterojunction to the 0.65 mA cm−2 at 1.23 V versus Ag/Ag-Cl electrode.
Table 2 provides some other research activities on the water-splitting application of the g-C3N4–metal oxide-based composites.

6.1.2. CO2 Reduction

CO2 emission is one of the leading environmental problems causing by fossil fuel consumption and results in a temperature rise of the earth’s surface. Photocatalytic CO2 reduction is a green method to deal with this problem for two reasons [111]. Not only CO2 reduction reduces the CO2 emission, but it also solves the future energy demands by producing energy fuels such as CH4, CH3OH, etc. Photocatalytic CO2 conversion is largely achieved by different metal oxide-g-C3N4 based systems, as they have desirable band edges positions. Equations (4)–(7) are chemical reactions for CO2 conversion to other solar fuels.
CO2 (g) + 2 H+ + 2 e → CO +H2O   E° = −0.53 eV vs. NHE at pH 7
CO2 (g) + 2 H+ + 2 e → HCOOH   E° = −0.61 eV vs. NHE at pH 7
CO2 (g) + 6 H+ + 6 e → CH3OH + H2O   E° = −0.38 eV vs. NHE at pH 7
CO2 (g) + 4 H+ + 4 e → HCHO + H2O   E° = −0.48 eV vs. NHE at pH 7
ZnO and TiO2 are widely used as the g-C3N4-based composites for CO2 conversion [429,430]. Wang et al. designed a photocatalyst comprising TiO2 and g-C3N4 using ball milling and calcination. The heterostructure between TiO2 and C3N4 leads to a low charge recombination rate, and high separation, resulting in the high CH4 and CO evolution yields of 72.2 and 56.2 μmol g−1 are obtained [431]. In another research, Nb-TiO2/g-C3N4 Z-scheme heterojunctions were investigated and showed that the 50Nb-TiO2/50g-C3N4 composition was the best photocatalysts with high carrier separation ability for the reduction of CO2 [432]. The existence of electrons and holes in the CB of the g-C3N4 and VB of Nb-TiO2, respectively, makes the Nb-TiO2/g-C3N4 system a potential candidate for reducing of CO2 into CH4 and CO and HCOOH. Guo and coworkers were thermally deposited g-C3N4 onto the porous ZnO nanosheets by two-step calcination and demonstrated that the ZnO porous nanosheets @ g-C3N4-0.4 showed the highest CO2 conversion efficiency [433]. Not only this composite suppressed the photoinduced electron recombination and facilitated the carrier transfer, but also the CO2 chemosorption increased in this composite since the increasing defect vacancies formed on the porous ZnO nanosheets. Shen et al. demonstrated 3-ZnO/g-C3N4 (3 is the mass ratio of ZnO) has a high photocatalytic activity for CO2 reduction to CO and CH4 [434]. This experiment showed an insignificant decrease in photocatalytic activities, which indicates ZnO-g-C3N4 had high photocatalytic stability. A similar analysis has been performed on the hydrocarbon generation rate with hollow g-C3N4, hollow CeO2, which is shown in Figure 18a [435]. The fast kinetics of CO reduction results in a higher CO evolution rate compared to the CH4 and CH3OH. The highest yield obtains for g-C3N4@ 49.7 wt % CeO2. Increasing amounts of CeO2 in the g-C3N4@CeO2 composites will decrease the PL intensity at about 460 nm, meaning that the lower electrons and holes recombination and enhanced charge separation (Figure 18b) [435].
Other researchers also work on the CO2 reduction of g-C3N4–metal oxide-based photocatalysts, which is listed in Table 3.

6.1.3. Photodegradation of Organic Pollutants

The development of an increased number of dye-related industries such as textile, food, and furniture manufacturing leads to severe environmental problems [445,446]. In addition to the negative aesthetic impact on water sources, the chemical oxygen demand (COD) in wastewater will be increased in the presence of organic dyes. Various methods, such as coagulation, adsorption, and membrane separation, have been used to eliminate organic dye from effluents, which only reclaim organic dyes from the wastewater liquid phase to the solid phase, creating secondary pollutants in the environment. These techniques are also a significant threat to living organisms. As a result, a metal oxide semiconductor has been widely used for the degradation of organic dyes. Photocatalysis in which most of the metal oxide can eliminate organic dyes by degradation and transfection them into particles, using solar energy for activation of the reaction. The related equations for dye degradation of the metal oxide-g-C3N4 are shown in Equations (8)–(11) [447]:
(g-C3N4+Metal Oxide) + hυ → eCB + hVB+
O2 + 2eCB + 2H+ → OH + OH
hVB+ + H2O → H+ + OH
Organic pollutant + OH → CO2 + H2O
Researchers have suggested several metal oxide semiconductors that can be used in the g-C3N4-based heterojunctions. There are increasing numbers of studies showing the degradation capability of TiO2, ZnO, WO3, Bi2O3, CeO2, etc. [100,101,113,116,448,449,450,451]. Zada et al. investigated the photodegrading of 2,4-dichlorophenol (2,4-DCP) and bisphenol A (BPA) over Au-(TiO2/g-C3N4) nanocomposites. As mentioned above, the excellent photocatalytic of the nanocomposites containing Au is mainly due to the SPR of decorated Au [131]. The structure revealed 46% and 37% for 2,4-DCP and BPA degradation, which is 5.11 and 3.1 times larger than the bulk g-C3N4 in water under visible-light irradiation, respectively [131]. In another research, the synthesized ZnO@g-C3N4 exhibited an enhanced photocatalytic activity to degrade tetracycline (TC) under visible-light irradiation, which is 2.77 and 1.51 fold more than the photocatalytic ability of pure g-C3N4 and ZnO [452]. The improved degradation was due to the enhanced transference of charge carriers and reduced charge recombination in the presence of the generated reactive oxygen species (ROS). The pharmaceuticals contaminants have hazardous impacts on human health and environmental biodiversity. Zhu and coworkers fabricated WO3-g-C3N4 composites for the photocatalytic degradation of one of the most well-known antibiotics, sulfamethoxazole (SMX), under visible light irradiation [453]. It is also revealed that the presence of RGO besides WO3-g-C3N4 heterojunctions promoted the degradation rate of ciprofloxacin (CIP) nearly twice as compared to the WO3-g-C3N4 structure [454]. In addition, H2O2 can raise the photocatalytic activities by the hydroxyl radicals’ productions from the degradation of a natural organic matter up to 71% for 5 h [455]. Shafawi et al. prepared Bi2O3 particles decorated on porous g-C3N4 sheets by impregnation method [456]. 1 g/L of the composite containing g-C3N4 with 9 wt % Bi2O3 at 10 ppm reactive black 5 (RB 5) at pH = 5.7 demonstrated 84% degradation efficiency under UV-vis light for 120 min. The synergistic effects between g-C3N4 and CeO2 provides higher catalytic activities compared to the bare ones [457,458,459]. The catalytic effects of the g-C3N4/CeO2 composite, bare g-C3N4, and CeO2 on the thermal decomposition of ammonium perchlorate (AP) were analyzed by using TGA and DTA characterization [460]. Two weight-loss regions from 25 °C to 500 °C, which were similar to the weight loss steps of AP in the absence of catalyst, were observed in Figure 19a,b. The weight loss decomposition temperatures of AP in the presence of the pure g-C3N4, CeO2, and g-C3N4/CeO2 were 53.6 °C, 47.6 °C and 74.6 °C, respectively (Figure 19a). It is also noticeable from Figure 19b that the AP thermal decomposition rate of g-C3N4/CeO2 nanocomposites was higher than that of CeO2 and g-C3N4. This heterojunction also shows a highly efficient for 2,4-dichlorophenol degradation [460].
Methylene Blue (MB), Methylene Orange (MO), and Rhodamine B (RhB) are the most commonly stable dyes in water at room temperature [461,462,463,464,465,466,467,468]. MB and RhB are toxic dyes that their high concentration can be so harmful to human and marine animal health. In accordance with the high resistivity of the MB and RhB in different environmental conditions, wastewater treatment is an urgent issue. Therefore, it is crucial to provide an effective and low-cost system to eliminate MB and RhB from sewage. As mentioned above, 2D/1D g-C3N4/ZnO nanocomposites reveal high stability, which retains the initial activity after repeated cycles [469]. The enhanced photocatalytic activity of g-C3N4-ZnO is likely due to the synergistic effects of photon acquisition and direct contact between organic dyes and photocatalyst [470]. To be more specific, the RhB dye degradation efficiency of this composite is 99% and 95% after one and three cycles under sunlight irradiation, respectively [471]. It is reported that the highest degradation efficiency for MB is obtained for the 30 wt % few-layer g-C3N4-ZnO nanocomposites is the best composition [472]. The higher amount of g-C3N4 can increase the electrons and holes recombination, leading to the decrease of photocatalytic activity. It is also noticeable that RhB degradation efficiency for this composite is about 2.1 times higher than that of pristine ZnO [473]. Apart from g-C3N4-ZnO heterojunction, g-C3N4-TiO2 heterojunction has high capability in the RhB degradation [474,475,476,477]. In these composites, O2 played a major role while h+ played a minor role [478,479,480]. Li et al. prepared Ti3+ self-doped TiO2 nanoparticles/g-C3N4 heterojunctions and demonstrated that the Ti3+ and O defects improve the conductivity and light absorption range to the visible wavelength region. Besides, the photocatalytic activities for environmental purification of organic compounds (MB) of the Ti3+ self-doped TiO2/g-C3N4 nanostructures were improved under visible light irradiation remarkably [481]. It is proven that the photocatalytic activity of 2 wt % g-C3N4-TiO2 improved by 70% compared to the bare TiO2 [118]. The RhB degradation kinetic constant of g-C3N4-TiO2 heterojunction is 9 times and 25 times higher than g-C3N4 and TiO2, respectively [482]. The AgPO4/g-C3N4 was used for the degradation of organic pollutants. Z-scheme carbon nitride with AgPO4 and Ag nanoparticles was used for the degradation of RhB. The presence of Ag can improve visible light absorption. The prepared hybrid structure can improve the electrons holes separation and efficiency [483]. Other metal oxides can also be used in the g-C3N4 based heterojunction for organic dye degradation [398,484,485,486,487]. Bi2O3/g-C3N4 heterojunctions are the most commonly used composites for organic dye degradation [488,489]. Fan et al. mentioned that the OH radicals played the crucial roles during photocatalytic degradation of MB in 0.5- Bi2O3/g-C3N4 (0.5 is the mass of Bi2O3 in the synthetic process) Z-scheme heterojunction, while O2, h+, e radicals have less contribution to this activity [490]. Compared to the pure Bi2O3 and bare g-C3N4, the Bi2O3/g-C3N4 heterojunctions showed better catalytic activity with 100% MB degradation ability in 90 min [490]. The 2% Bi2O3/g-C3N4 composites showed a degradation rate constant of 0.040 min−1, which is 2.5 and 1.9 fold higher than that found for bulk and nitrogen vacant 2D g-C3N4 nanosheet, respectively [491]. This composition showed excellent photocatalytic activity with a methylene green degradation efficiency of 98.7% under visible light irradiation [448]. At 30 °C, 5% CeO2/g-C3N4 (5% is the molar ratios of the CeO2/g-C3N4 samples) photocatalyst showed the best efficiency for MB degrading under visible light irradiation with the constant rate of 1.2686 min−1, which is 7.8-fold higher than pure g-C3N4 [492]. Some ternary composites are also used for RhB and MB dye degradation [493]. SnO2−ZnO quantum dots anchored on g-C3N4 nanosheets were demonstrated as a promising candidate for RhB degradation, which is 99% in 60 min under visible-light irradiation [494]. This hybrid also showed a promising potential for hydrogen production with the photocatalytic rate of 13.61 μmol g−1, which is 1.06 and 2.27 times higher than that of the binary ZnO/g-C3N4 hybrid and pristine g-C3N4. In similar work, ZnS quantum dots (ZNS)/SnO2/g-C3N4 ternary nanocomposites were synthesized via solid-state calcination [495]. The high bandgap of SnO2 and ZnS QDs leads to the reduction of the recombination rate and increases the separation of the generated carriers in the g-C3N4 [495]. The lower conduction band edge position of the SnO2, which is lower than others compel electrons to accumulate in this state. The position of the electrons and holes helped in improving the photocatalytic performance and enhancing stability. As a result, the transferred electrons can act as a suitable reductant and react with adsorbed O2 to produce superoxide radicals (O2). On the other side, the holes in the valence band of g-C3N4 can directly oxidize the pollutants to degraded products since these holes have strong oxidizing power [495]. g-C3N4–metal oxide heterojunction is used to degrade different elements such as Hg, Cr (VI) [136,496,497], and different toxic and harmful materials such as atrazine (ATZ), chloramphenicol, ciprofloxacin (CIP), and etc. [197,454,498,499,500,501,502]. There are several research activities on the photodegradation applications of g-C3N4–metal oxide-based heterojunctions which are listed in Table 4.

6.2. Sensors

Some benefits, such as high sensitivity to analytes, rapid response to external stimulations, excellent fluorescence quenching abilities, light and electricity conversion properties, biocompatibility, and high stability make g-C3N4 nanosheet a promising candidate as a modified electrode for sensors to detect different analytes such as glucose, hydrogen peroxide, dopamine, etc. [6,549,550,551,552,553,554]. Metal oxide semiconductors/g-C3N4 composites are widely used as gas sensors [555]. As a result, the g-C3N4 loaded with metal oxides has also revealed new types of sensors to detect different kinds of materials [556]. General schematic description in Figure 20, shows different types of g-C3N4–metal oxide sensors.
Like other applications, g-C3N4–TiO2 based structures are one of the most used composites for sensing applications. Li et al. demonstrated that the photoelectrochemical TiO2/g-C3N4/CdS platform could be employed for the ultrasensitive determination of T4 polynucleotide kinase (T4 PNK) because this composite showed significant stability and reproducibility with high selectivity [557]. The photocurrent response is demonstrated in Figure 21A. To be more specific, “trace a” shows a low photocurrent at a bare electrode. The rapidly increased photocurrent is detected in “trace b” (from 10 to 30 s) due to decrease in the rate of electron-hole recombination and the high-efficiency absorption. In the TiO2/g-C3N4 modified-FTO in “trace c”, the photocurrent increases to 30.7 μA since the bandgap of the g-C3N4 reduces the charge carrier recombination and facilitates the electron transmission [557]. The photocurrent increases in “trace d” (TiO2/g-C3N4/CdS nanocomposite-modified electrode). The photocurrent increased to 80.0 μA after attaching DNA3, which can capture ssDNA2 to the electrode and then blocking 6-mercaptohexanol (MCH) (“trace e”). For comparison, the photocurrent of CdSe QDs with or without the bio-functionalization of DNA2 on the electrode matrix is 180 and 78 μA, respectively [557]. In addition, “trace f” shows that DNA2-CdSe QDs hybridized onto the surface, the photocurrent density increases to 180.0 μA since the presence of CdSe QDs improve the light absorption efficiency. To compare, CdSe QDs also functionalized the electrode matrix without DNA2, and the result showed that the photocurrent density is about 78.0 μA which is similar to the “trace e” and highly lower than “trace f”. As a result, the constructed platform shows a promising candidate for ultrasensitive detection of T4 PNK activity [557]. TiO2 with g-C3N4 is also used for sensitive detection of protein kinase A (PKA), which is an enzyme that covalently decorates proteins with phosphate groups [558].
ZnO-g-C3N4 is another heterojunction used as an agent as UV-assisted gas sensors. It is shown that the ethanol (C2H5OH) sensing capability of ZnO-g-C3N4 is much higher when compared to the bare ZnO and g-C3N4. The ZnO with 8% g-C3N4 showed the best sensing performance than the others, which attributes to effective electrons and holes separation between g-C3N4 and ZnO and the UV-light catalytic effect in the room temperature [559]. The structure showed an excellent response of ethanol at room temperature, which was higher than the pure ZnO at the same condition. The arc radius of ZnO and g-C3N4/ZnO is shown in Figure 21B. The results show that adding g-C3N4 nanosheets to ZnO improves the charge transfer and reduces the electrical resistivity [559].
Sensing of NO2 and CH4 by ZnO-g-C3N4 heterojunction have also been investigated [560,561]. Hydrogen sulfide (H2S) is a corrosive and toxic gas, which can be generated from oil refining, mines, or petroleum fields, fuel cells, or food processing industries and may cause severe problems. As a result, eliminating H2S has always been a serious research topic. Zeng et al. used α-Fe2O3/g-C3N4 for H2S detection and showed that 5.97% α-Fe2O3 with g-C3N4 provides the best cataluminescence response [562]. The prepared α-Fe2O3/g-C3N4 composite can be used as an H2S gas sensor in environmental monitoring, oil refining, food processing, and so forth. Researchers have shown that g-C3N4-SnO2 composites can also be used as sensors for different kinds of compounds. Specifically, Zhang et al. proved that the presence of g-C3N4 beside SnO2 enhances the gas sensitivity and selectivity of SnO2 to acetic acid vapor with the reduced temperature from 230 °C to 185 °C [563]. The limit of acetic acid vapor detection is 0.1 ppm; however, the long-term stability of the prepared composite is low and should be improved. In-SnO2 loaded cubic mesoporous g-C3N4 will be a new method and structure to design efficient humidity sensors for monitoring indoor climatic conditions [564]. Table 5 lists the applications of g-C3N4–metal oxide heterojunction used as sensor material.

6.3. Bacterial Disinfection

According to different sizes, shapes, and structures of bacteria, viruses, microbes, and microalgae, the g-C3N4 has the ability to eliminate them under ultraviolet or visible light irradiation changes. Fabricating the g-C3N4 with different metal oxides is important for effective composites for water antimicrobial disinfection and microbial control. Biohazards are widely present in wastewater and contaminated water containing a variety of viruses, bacteria, fungi, etc., causing health issues in humans [30]. The pioneering work by Matsunaga et al. showed that TiO2 could help in inactivating bacteria such as Escherichia coli, Lactobacillus acidophilus, and Saccharomyces cerevisiae under UV light [580]. TiO2 suffers from low absorption of solar energy and this is due to the wide bandgap of TiO2, which is widely discussed in the previous parts. One method to increase TiO2 performance is to combine it with different semiconductors. Li et al. prepared g-C3N4-TiO2 using a facile hydrothermal-calcination approach with high photocatalytic bacterial inactivation ability [114]. They demonstrated that this composite could facilitate water disinfection, especially in hospital wastewaters with highly concentrated pathogenic microorganisms, using visible light. In another work, researchers investigated the excellent efficiency of this composite on the removal of the microcystis aeruginosa and Microcystin-LR [581].
In order to improve the photocatalytic activity, researchers are more likely to dope metal elements such as Ag and Cr in the g-C3N4–metal oxide structure, which can perform as an electron sink to separate the carriers. Doping these elements not only can modify the band edge positions of these structures but also, due to antimicrobial and bacterial properties of these materials, they are widely used in the g-C3N4–metal oxide-based photocatalysts [102,189]. g-C3N4/Cr-ZnO nanocomposites with superior antibacterial activity against Gram-negative (Escherichia coli) and Gram-positive (Bacillus subtilis, Staphylococcus aureus, and Streptococcus salivarius) were investigated [172]. In this research, the 60%g-C3N4/5%Cr-ZnO nanocomposite performs the highest antibacterial activity. Fe-SnO2/g-C3N4 revealed a promising sterilization performance for Escherichia coli and Staphylococcus aureus under sunlight, near-ultraviolet light, and daylight lamp [582]. The sterilization performance of this structure is mostly deserved under a daylight lamp. The magnetic silver-iron oxide nanoparticles decorated graphitic carbon nitride nanosheets showed antibacterial performance against E. coli bacteria [583].
Zhao and coworkers illustrated that g-C3N4/ZnO/cellulose (CNZCel) shows the high thermal stability of the composite and photo-excited carriers separation efficiency and decreases the recombination rate [202]. Besides, the nanocomposite revealed excellent antibacterial activity against Escherichia coli (E. coli) and Gram-positive bacteria Staphylococcus aureus (S. aureus). The presence of ZnO in the structure can efficiently enhance the antibacterial activities against E. coli and S. aureus. Compared to the other materials, these ternary composites demonstrated better antibacterial performance (Figure 22a,b). The effect of the g-C3N4 content on the antibacterial property against E. coli and S. aureus in this composite is also investigated, which is shown in Figure 22c,d [202]. The mechanism of antibacterial activity generates electron-hole pairs under illumination, which can provide enormous reactive oxygen species (ROS), such as O2, OH2 and H2O2, needed for the inhibition of E. coli and S. aureus growth.
Table 6 mentioned some additional studies on the disinfection ability of g-C3N4–metal oxide-based nanomaterials.

6.4. Other Applications

Due to the specific properties of g-C3N4-based heterojunctions, these composites can be used in many other applications [150,247,585,592]. With reference to the high energy demand for electronic devices and vehicles, rechargeable lithium-ion batteries (LIBs) have captured enormous attention among many researchers [319,320,593,594,595]. Anode materials, which are a significant part of the LIBs, should have improved specific capacity, high stability. SnO2 is a new type of lithiophilic material with high specific capacity (1494 mAh g−1), low potential for Li+ insertion, increased number of sources, and etc. [320,593,596,597,598,599,600,601,602]; however, it suffers from significant volume expansion (∼300%) during the charge-discharge cycling, leading to fast capacity fading. SnO2 nanosheets with 20–25 nm thickness dispersed in the g-C3N4 showing the potential lithium storage for LIBs. SnO2@C3N4 nanocomposites can also be prepared by a scalable solid-state reaction [319]. Tran et al. used the hydrothermal method to grow SnO2 onto the graphite oxide/g-C3N4 [320]. This composite showed an excellent reversible capacity and cycling performance for lithium storage, which may attribute to the existence of g-C3N4 or graphite oxide-g-C3N4. SnO2@g-C3N4 based nanocomposites provide a suitable substitute for next-generation high-power and low-cost LIBs [319]. Zn2GeO4 nanoparticles demonstrate high capacity and act as spacers to prevent the g-C3N4 sheets from stacking, leading to the expanded interlayer and exposed vacancies for higher Li-ion storage. Besides, g-C3N4 layers, in turn, reduce the expansion of the particles and provide more stable solid electrolyte interphase, results in highly reversible lithium storage capacity, which is 1370 mA h g−1 at 200 mA g−1 after 140 cycles with a significant rate capability of 950 mA h g−1 at 2000 mA g−1 [595]. The SnO2 nanosheets with g-C3N4 can enhance lithium storage capabilities and cycling performance [593]. Other g-C3N4–metal oxide nanocomposites can also be used in the supercapacitor that can be used with batteries to mitigate the power delivery problems associated with batteries [596,597,598]. The bare g-C3N4, g-C3N4/CuO, g-C3N4/Co3O4 electrode-based device exhibited a specific capacitance of 72 F g−1, 95 F g−1, and 201 F g−1, respectively [599]. Besides, the energy density of g-C3N4/CuO and g-C3N4/Co3O4 at the constant power density of 1 kW·kg−1 are 13.2 W·h·kg−1, and 27.9 W·h·kg−1, respectively. Moreover, the excellent theoretical capacities of V2O5 for the next generation supercapacitors makes the porous g-C3N4@V2O5 good candidate with a high specific capacity of about 457 Fg−1 at 0.5 Ag−1 with high cycling performance (~84% after 500 cycles) [600].
Photocatalytic nitrogen fixation, which is a clean and sustainable method for the production of NH3, is another application of g-C3N4-based materials. However, insignificant surface-active sites, poor stability, and high recombination rate of carriers have restricted the efficiency of g-C3N4 for nitrogen fixation activities. The carrier separation should be enhanced by doping with other elements or compositing with various metal oxides. Finally, nitrogen gas should completely adsorb on the photocatalyst for the reduction steps [603]. It is shown that the cyano group (–C☰NC) in cyano group modified g-C3N4 enhance the photocatalytic applications for N2 fixation up to 128 times [604].
As we discussed earlier, iodine and sulfur are atoms that can improve the carriers’ activities in the bulk g-C3N4 [605]. To be more specific, ultrathin sulfur-doped g-C3N4 porous nanosheets revealed a superior photocatalytic nitrogen fixation rate with 5.99 mM·h−1·g−1 for 4 h under simulated sunlight irradiation [606]. Zhang et al. demonstrated that B atoms change the band structures and WO3 enhances the photocatalytic activities. The quantum efficiency (QE) of the prepared composites is 0.71% for nitrogen fixation at 400 nm with a yield of 450.94 μmol g−1 h−1 under visible light [606]. In another work, Liu and coworkers mentioned that the phosphorus-doped 1 T-MoS2 as co-catalyst decorated nitrogen-doped g-C3N4 nanosheets speed up the N2 reduction rate to 689.76 μmol L−1 g−1·h−1 in deionized water under simulated sunlight irradiation, which is 2.59, 1..65, 1.47, and 1.30 times higher than that of pure g-C3N4, 1 T-MoS2@g-C3N4, 1 T-MoS2@N doped g-C3N4, and P doped 1 T-MoS2@g-C3N4, respectively. It is noteworthy to note that MoS2 mainly includes semiconductive 2H phase and metal octahedral 1 T phase with promising conductivity and suitable photocatalysis since this phase provides a large number of active sites on the base and edge [607]. Researchers also used metal oxides with g-C3N4-based composites to improve nitrogen fixation productivity. TiO2/SrTiO3/g-C3N4 ternary heterojunction nanofibers demonstrated N2 fixation value under-stimulated sunlight irradiation is 2192 μmol g−1 h−1 L−1, which is 1.9 and 3.3 times better than those of TiO2/g-C3N4 nanofibers and SrTiO3/g-C3N4 nanofibers, respectively [605,608]. Oxidized g-C3N4 can also enhance the chemo selective and unselective oxidative processes, utilized in organic reactions to provide synthons required in several value-added preparations [609].
It is interesting to mention that g-C3N4–metal oxide-based heterojunction can be widely used to detect of different biological materials, gasses, heavy metals, organic and inorganic materials [547,610]. g-C3N4–metal oxide-based composites can also be widely used for other applications, especially supercapacitors and desulfurization [107,611,612,613].

7. Conclusions, Limitations and Challenges

To fix inadequate light-harvesting of the bare g-C3N4, many researchers used doping elements to improve the efficiency, while combining g-C3N4 with other materials, such as metal oxides are another approach to deal with this issue. The investigation of g-C3N4–metal oxide-based heterojunctions with engineered bandgaps, and modified surface to enhance the abortion spectrum towards the visible-light region, reduction of the charge carrier recombination, and increasing the surface adsorption and reaction are among the main scopes of investigations of g-C3N4 based materials. Among all types of heterojunctions mentioned in the review paper, almost all g-C3N4–metal oxide-based nanocomposites have a Type II or a Z-scheme system, which have been proven to be outstanding for different applications. The applications highlighted included water splitting, CO2 reduction, photodegradation of organic pollutants, nitrogen fixation, catalysis, sensing, bacterial disinfection, energy storage, etc.
Several types of carbon-based nanomaterials such as carbon-based quantum dots (CDs) (carbon quantum dots (CQDs) and graphene quantum dots (GQDs)), which are discussed in our previous work, can be used as sensors [214]. One kind of carbon-based nanomaterials is g-C3N4, which is widely used to detect different cells in human biological fluids, gasses, humidity, heavy metals, etc. Nevertheless, these applications also can be enormously challenging and give plenty of room for research and development, and there are great demands for fast and accurate diagnostic methods for both clinical and commercial applications. As a result, it is suggested to develop systems for clinical use to detect different diseases outside the body, especially the newly emerging one, COVID- 19, mainly due to its high chemical stability and excellent efficiency in absorbing the light and promising photocatalytic performance. In addition, the detection of different heavy metals and ions is one of the most crucial factors for contaminant removal and water treatment. Since g-C3N4–metal oxide photocatalysts have exhibited the superior potential to detect different materials, it seems critical to extending investigations on the efficiency of these heterojunctions for detection of Cu2+, Hg2+, Cr (VI), and Ag+.
The need for using solar energy has increased among many researchers, as energy resources are going to be depleted. Researchers require suitable semiconductors for different catalytic reactions, such as H2 and O2 generation by water splitting, CO2 reduction to hydrocarbon fuels, etc. In order to have a future practical application in this area, we have to resolve some issues about the conversion of solar energy to fuel. The cocatalysis and quantum efficiency are among the most important properties for H2O splitting and CO2 reduction, which for the g-C3N4–metal oxide composites, researchers do not have enough knowledge. The enhanced photocatalytic activity of the mentioned nanocomposites in this review under visible light encourages researchers to use these heterojunctions in antibacterial and antiviral applications. Inactivation originates from direct hole oxidation and O2, OH, and H2O2, which are produced in a reductive way, and dependent on the type of microorganisms. Because several research works have been conducted on water disinfection and microbial control in the laboratories, it is necessary to have more practical exploration on full-scale water treatment.

Author Contributions

A.A.: Investigation, Reviewing papers, Writing the first draft. K.G.: Conceptualization, Supervision, Writing-Reviewing and Editing, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research received was funded by NSERC Discovery Grant of Dr. Ghandi to support the PhD of Amirhossein Alaghmandfard at University of Guelph.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Goettmann, F.; Fischer, A.; Antonietti, M.; Thomas, A. Metal-Free Catalysis of Sustainable Friedel–Crafts Reactions: Direct Activation of Benzene by Carbon Nitrides to Avoid the Use of Metal Chlorides and Halogenated Compounds. Chem. Commun. 2006, 43, 4530–4532. [Google Scholar] [CrossRef] [PubMed]
  2. Ong, W.-J.; Tan, L.-L.; Ng, Y.H.; Yong, S.-T.; Chai, S.-P. Graphitic Carbon Nitride (g-C3N4)-Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer to Achieving Sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, X.; Ma, R.; Zhuang, L.; Hu, B.; Chen, J.; Liu, X.; Wang, X. Recent Developments of Doped g-C3N4 Photocatalysts for the Degradation of Organic Pollutants. Crit. Rev. Environ. Sci. Technol. 2020, 51, 751–790. [Google Scholar] [CrossRef]
  4. Taha, M.M.; Ghanem, L.G.; Hamza, M.A.; Allam, N.K. Highly Stable Supercapacitor Devices Based on Three-Dimensional Bioderived Carbon Encapsulated g-C3N4 Nanosheets. ACS Appl. Energy Mater. 2021, 4, 10344–10355. [Google Scholar] [CrossRef]
  5. Jiang, H.; Li, Y.; Wang, D.; Hong, X.; Liang, B. Recent Advances in Heteroatom Doped Graphitic Carbon Nitride (g-C3N4) and g-C3N4/Metal Oxide Composite Photocatalysts. Curr. Org. Chem. 2020, 24, 673–693. [Google Scholar] [CrossRef]
  6. Wang, L.; Wang, C.; Hu, X.; Xue, H.; Pang, H. Metal/Graphitic Carbon Nitride Composites: Synthesis, Structures, and Applications. Chem. Asian J. 2016, 11, 3305–3328. [Google Scholar] [CrossRef] [PubMed]
  7. Ren, Y.; Zeng, D.; Ong, W.-J. Interfacial Engineering of Graphitic Carbon Nitride (g-C3N4)-Based Metal Sulfide Heterojunction Photocatalysts for Energy Conversion: A Review. Chin. J. Catal. 2019, 40, 289–319. [Google Scholar] [CrossRef]
  8. Ismael, M. A Review on Graphitic Carbon Nitride (g-C3N4) Based Nanocomposites: Synthesis, Categories, and Their Application in Photocatalysis. J. Alloys Compd. 2020, 846, 156446. [Google Scholar] [CrossRef]
  9. Inagaki, M.; Tsumura, T.; Kinumoto, T.; Toyoda, M. Graphitic Carbon Nitrides (g-C3N4) with Comparative Discussion to Carbon Materials. Carbon 2019, 141, 580–607. [Google Scholar] [CrossRef]
  10. Chu, Y.-C.; Lin, T.-J.; Lin, Y.-R.; Chiu, W.-L.; Nguyen, B.-S.; Hu, C. Influence of P, S, O-Doping on g-C3N4 for Hydrogel Formation and Photocatalysis: An Experimental and Theoretical Study. Carbon 2020, 169, 338–348. [Google Scholar] [CrossRef]
  11. Wei, T.; Xu, J.; Kan, C.; Zhang, L.; Zhu, X. Au Tailored on g-C3N4/TiO2 Heterostructure for Enhanced Photocatalytic Performance. J. Alloys Compd. 2022, 894, 162338. [Google Scholar] [CrossRef]
  12. Gotipamul, P.P.; Vattikondala, G.; Rajan, K.D.; Khanna, S.; Rathinam, M.; Chidambaram, S. Impact of Piezoelectric Effect on the Heterogeneous Visible Photocatalysis of g-C3N4/Ag/ZnO Tricomponent. Chemosphere 2022, 287, 132298. [Google Scholar] [CrossRef]
  13. Fu, J.; Yu, J.; Jiang, C.; Cheng, B. g-C3N4-Based Heterostructured Photocatalysts. Adv. Energy Mater. 2018, 8, 1701503. [Google Scholar] [CrossRef]
  14. Zou, W.; Deng, B.; Hu, X.; Zhou, Y.; Pu, Y.; Yu, S.; Ma, K.; Sun, J.; Wan, H.; Dong, L. Crystal-Plane-Dependent Metal Oxide-Support Interaction in CeO2/g-C3N4 for Photocatalytic Hydrogen Evolution. Appl. Catal. B Environ. 2018, 238, 111–118. [Google Scholar] [CrossRef]
  15. Li, J.; Yuan, H.; Zhu, Z. Improved Photoelectrochemical Performance of Z-Scheme g-C3N4/Bi2O3/BiPO4 Heterostructure and Degradation Property. Appl. Surf. Sci. 2016, 385, 34–41. [Google Scholar] [CrossRef]
  16. Liu, C.; Qiu, Y.; Zhang, J.; Liang, Q.; Mitsuzaki, N.; Chen, Z. Construction of CdS Quantum Dots Modified g-C3N4/ZnO Heterostructured Photoanode for Efficient Photoelectrochemical Water Splitting. J. Photochem. Photobiol. A Chem. 2019, 371, 109–117. [Google Scholar] [CrossRef]
  17. Sun, Q.; Lv, K.; Zhang, Z.; Li, M.; Li, B. Effect of Contact Interface between TiO2 and g-C3N4 on the Photoreactivity of g-C3N4/TiO2 Photocatalyst:(0 0 1) vs (1 0 1) Facets of TiO2. Appl. Catal. B Environ. 2015, 164, 420–427. [Google Scholar]
  18. Liang, B.; Han, D.; Sun, C.; Zhang, W.; Qin, Q. Synthesis of SnO/g-C3N4 Visible Light Driven Photocatalysts via Grinding Assisted Ultrasonic Route. Ceram. Int. 2018, 44, 7315–7318. [Google Scholar] [CrossRef]
  19. Dai, L.; Chang, D.W.; Baek, J.; Lu, W. Carbon Nanomaterials for Advanced Energy Conversion and Storage. small 2012, 8, 1130–1166. [Google Scholar] [CrossRef]
  20. Gong, Y.; Wang, J.; Wei, Z.; Zhang, P.; Li, H.; Wang, Y. Combination of Carbon Nitride and Carbon Nanotubes: Synergistic Catalysts for Energy Conversion. ChemSusChem 2014, 7, 2303–2309. [Google Scholar] [CrossRef]
  21. Mao, Z.; Chen, J.; Yang, Y.; Wang, D.; Bie, L.; Fahlman, B.D. Novel g-C3N4/CoO Nanocomposites with Significantly Enhanced Visible-Light Photocatalytic Activity for H2 Evolution. ACS Appl. Mater. Interfaces 2017, 9, 12427–12435. [Google Scholar] [CrossRef]
  22. Zou, Y.; Shi, J.-W.; Ma, D.; Fan, Z.; Lu, L.; Niu, C. In Situ Synthesis of C-Doped TiO2@ g-C3N4 Core-Shell Hollow Nanospheres with Enhanced Visible-Light Photocatalytic Activity for H2 Evolution. Chem. Eng. J. 2017, 322, 435–444. [Google Scholar] [CrossRef] [Green Version]
  23. Hasanvandian, F.; Moradi, M.; Samani, S.A.; Kakavandi, B.; Setayesh, S.R.; Noorisepehr, M. Effective Promotion of g-C3N4 Photocatalytic Performance via Surface Oxygen Vacancy and Coupling with Bismuth-Based Semiconductors towards Antibiotics Degradation. Chemosphere 2022, 287, 132273. [Google Scholar] [CrossRef]
  24. Zhang, C.; Qin, D.; Zhou, Y.; Qin, F.; Wang, H.; Wang, W.; Yang, Y.; Zeng, G. Dual Optimization Approach to Mo Single Atom Dispersed g-C3N4 Photocatalyst: Morphology and Defect Evolution. Appl. Catal. B Environ. 2022, 303, 120904. [Google Scholar] [CrossRef]
  25. Zhang, S.; Gu, P.; Ma, R.; Luo, C.; Wen, T.; Zhao, G.; Cheng, W.; Wang, X. Recent Developments in Fabrication and Structure Regulation of Visible-Light-Driven g-C3N4-Based Photocatalysts towards Water Purification: A Critical Review. Catal. Today 2019, 335, 65–77. [Google Scholar] [CrossRef]
  26. Thomas, A.; Fischer, A.; Goettmann, F.; Antonietti, M.; Müller, J.-O.; Schlögl, R.; Carlsson, J.M. Graphitic Carbon Nitride Materials: Variation of Structure and Morphology and Their Use as Metal-Free Catalysts. J. Mater. Chem. 2008, 18, 4893–4908. [Google Scholar] [CrossRef] [Green Version]
  27. Ye, L.; Wang, D.; Chen, S. Fabrication and Enhanced Photoelectrochemical Performance of MoS2/S-Doped g-C3N4 Heterojunction Film. ACS Appl. Mater. Interfaces 2016, 8, 5280–5289. [Google Scholar] [CrossRef] [PubMed]
  28. Arumugam, M.; Tahir, M.; Praserthdam, P. Effect of Nonmetals (B, O, P, and S) Doped with Porous g-C3N4 for Improved Electron Transfer towards Photocatalytic CO2 Reduction with Water into CH4. Chemosphere 2022, 286, 131765. [Google Scholar] [CrossRef] [PubMed]
  29. Yang, S.; Sun, Q.; Han, W.; Shen, Y.; Ni, Z.; Zhang, S.; Chen, L.; Zhang, L.; Cao, J.; Zheng, H. A Simple and High Efficient Composite Based on g-C3N4 for Super Rapid Removal of Multiple Organic Dyes in Water under Sunlight. Catal. Sci. Technol. 2022, in press.
  30. Zhang, C.; Li, Y.; Shuai, D.; Shen, Y.; Xiong, W.; Wang, L. Graphitic Carbon Nitride (g-C3N4)-Based Photocatalysts for Water Disinfection and Microbial Control: A Review. Chemosphere 2019, 214, 462–479. [Google Scholar] [CrossRef] [PubMed]
  31. Yan, S.C.; Li, Z.S.; Zou, Z.G. Photodegradation Performance of g-C3N4 Fabricated by Directly Heating Melamine. Langmuir 2009, 25, 10397–10401. [Google Scholar] [CrossRef]
  32. Praus, P.; Svoboda, L.; Ritz, M.; Troppová, I.; Šihor, M.; Kočí, K. Graphitic Carbon Nitride: Synthesis, Characterization and Photocatalytic Decomposition of Nitrous Oxide. Mater. Chem. Phys. 2017, 193, 438–446. [Google Scholar] [CrossRef]
  33. Giannakopoulou, T.; Papailias, I.; Todorova, N.; Boukos, N.; Liu, Y.; Yu, J.; Trapalis, C. Tailoring the Energy Band Gap and Edges’ Potentials of g-C3N4/TiO2 Composite Photocatalysts for NOx Removal. Chem. Eng. J. 2017, 310, 571–580. [Google Scholar] [CrossRef]
  34. Tay, Q.; Kanhere, P.; Ng, C.F.; Chen, S.; Chakraborty, S.; Huan, A.C.H.; Sum, T.C.; Ahuja, R.; Chen, Z. Defect Engineered g-C3N4 for Efficient Visible Light Photocatalytic Hydrogen Production. Chem. Mater. 2015, 27, 4930–4933. [Google Scholar] [CrossRef]
  35. Mo, Z.; She, X.; Li, Y.; Liu, L.; Huang, L.; Chen, Z.; Zhang, Q.; Xu, H.; Li, H. Synthesis of g-C3N4 at Different Temperatures for Superior Visible/UV Photocatalytic Performance and Photoelectrochemical Sensing of MB Solution. RSC Adv. 2015, 5, 101552–101562. [Google Scholar] [CrossRef]
  36. Paul, D.R.; Sharma, R.; Nehra, S.P.; Sharma, A. Effect of Calcination Temperature, PH and Catalyst Loading on Photodegradation Efficiency of Urea Derived Graphitic Carbon Nitride towards Methylene Blue Dye Solution. RSC Adv. 2019, 9, 15381–15391. [Google Scholar] [CrossRef] [Green Version]
  37. Fan, X.; Xing, Z.; Shu, Z.; Zhang, L.; Wang, L.; Shi, J. Improved Photocatalytic Activity of g-C3N4 Derived from Cyanamide–Urea Solution. RSC Adv. 2015, 5, 8323–8328. [Google Scholar] [CrossRef] [Green Version]
  38. Liu, H.; Yu, D.; Sun, T.; Du, H.; Jiang, W.; Muhammad, Y.; Huang, L. Fabrication of Surface Alkalinized g-C3N4 and TiO2 Composite for the Synergistic Adsorption-Photocatalytic Degradation of Methylene Blue. Appl. Surf. Sci. 2019, 473, 855–863. [Google Scholar] [CrossRef]
  39. Li, F.; Zhao, R.; Yang, B.; Wang, W.; Liu, Y.; Gao, J.; Gong, Y. Facial Synthesis of Dandelion-like g-C3N4/Ag with High Performance of Photocatalytic Hydrogen Production. Int. J. Hydrogen Energy 2019, 44, 30185–30195. [Google Scholar] [CrossRef]
  40. Hu, C.; Hung, W.-Z.; Wang, M.-S.; Lu, P.-J. Phosphorus and Sulfur Codoped g-C3N4 as an Efficient Metal-Free Photocatalyst. Carbon 2018, 127, 374–383. [Google Scholar] [CrossRef]
  41. Chen, Y.; Wang, B.; Lin, S.; Zhang, Y.; Wang, X. Activation of N→ Π* Transitions in Two-Dimensional Conjugated Polymers for Visible Light Photocatalysis. J. Phys. Chem. C 2014, 118, 29981–29989. [Google Scholar] [CrossRef]
  42. Dias, E.M.; Christoforidis, K.C.; Francas, L.; Petit, C. Tuning Thermally Treated Graphitic Carbon Nitride for H2 Evolution and CO2 Photoreduction: The Effects of Material Properties and Mid-Gap States. ACS Appl. Energy Mater. 2018, 1, 6524–6534. [Google Scholar] [CrossRef]
  43. Ding, Z.; Chen, X.; Antonietti, M.; Wang, X. Synthesis of Transition Metal-modified Carbon Nitride Polymers for Selective Hydrocarbon Oxidation. ChemSusChem 2011, 4, 274–281. [Google Scholar] [CrossRef]
  44. Yuan, Y.; Zhang, L.; Xing, J.; Utama, M.I.B.; Lu, X.; Du, K.; Li, Y.; Hu, X.; Wang, S.; Genç, A. High-Yield Synthesis and Optical Properties of g-C3N4. Nanoscale 2015, 7, 12343–12350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Wang, X.; Maeda, K.; Chen, X.; Takanabe, K.; Domen, K.; Hou, Y.; Fu, X.; Antonietti, M. Polymer Semiconductors for Artificial Photosynthesis: Hydrogen Evolution by Mesoporous Graphitic Carbon Nitride with Visible Light. J. Am. Chem. Soc. 2009, 131, 1680–1681. [Google Scholar] [CrossRef] [PubMed]
  46. Creighton, J.R.; Ho, P. Introduction to Chemical Vapor Deposition (CVD). Chem. Vap. Depos. 2001, 2, 1–22. [Google Scholar]
  47. Xu, Z.; Guan, L.; Li, H.; Sun, J.; Ying, Z.; Wu, J.; Xu, N. Structure Transition Mechanism of Single-Crystalline Silicon, g-C3N4, and Diamond Nanocone Arrays Synthesized by Plasma Sputtering Reaction Deposition. J. Phys. Chem. C 2015, 119, 29062–29070. [Google Scholar] [CrossRef]
  48. Zhang, J.; Wang, Y.; Jin, J.; Zhang, J.; Lin, Z.; Huang, F.; Yu, J. Efficient Visible-Light Photocatalytic Hydrogen Evolution and Enhanced Photostability of Core/Shell CdS/g-C3N4 Nanowires. ACS Appl. Mater. Interfaces 2013, 5, 10317–10324. [Google Scholar] [CrossRef]
  49. Sano, T.; Tsutsui, S.; Koike, K.; Hirakawa, T.; Teramoto, Y.; Negishi, N.; Takeuchi, K. Activation of Graphitic Carbon Nitride (g-C3N4) by Alkaline Hydrothermal Treatment for Photocatalytic NO Oxidation in Gas Phase. J. Mater. Chem. A 2013, 1, 6489–6496. [Google Scholar] [CrossRef]
  50. Cui, Y.; Ding, Z.; Fu, X.; Wang, X. Construction of Conjugated Carbon Nitride Nanoarchitectures in Solution at Low Temperatures for Photoredox Catalysis. Angew. Chemie 2012, 124, 11984–11988. [Google Scholar] [CrossRef]
  51. Muniandy, L.; Adam, F.; Mohamed, A.R.; Iqbal, A.; Rahman, N.R.A. Cu2+ Coordinated Graphitic Carbon Nitride (Cu-g-C3N4) Nanosheets from Melamine for the Liquid Phase Hydroxylation of Benzene and VOCs. Appl. Surf. Sci. 2017, 398, 43–55. [Google Scholar] [CrossRef]
  52. Hu, C.; Chu, Y.-C.; Lin, Y.-R.; Yang, H.-C.; Wang, K.-H. Photocatalytic Dye and Cr (VI) Degradation Using a Metal-Free Polymeric g-C3N4 Synthesized from Solvent-Treated Urea. Polymers. 2019, 11, 182. [Google Scholar] [CrossRef] [Green Version]
  53. Liang, L.; Cong, Y.; Wang, F.; Yao, L.; Shi, L. Hydrothermal Pre-Treatment Induced Cyanamide to Prepare Porous g-C3N4 with Boosted Photocatalytic Performance. Diam. Relat. Mater. 2019, 98, 107499. [Google Scholar] [CrossRef]
  54. Cao, J.; Fan, H.; Wang, C.; Ma, J.; Dong, G.; Zhang, M. Facile Synthesis of Carbon Self-Doped g-C3N4 for Enhanced Photocatalytic Hydrogen Evolution. Ceram. Int. 2020, 46, 7888–7895. [Google Scholar] [CrossRef]
  55. Hong, Y.; Liu, E.; Shi, J.; Lin, X.; Sheng, L.; Zhang, M.; Wang, L.; Chen, J. A Direct One-Step Synthesis of Ultrathin g-C3N4 Nanosheets from Thiourea for Boosting Solar Photocatalytic H2 Evolution. Int. J. Hydrogen Energy 2019, 44, 7194–7204. [Google Scholar] [CrossRef]
  56. Wehrstedt, K.-D.; Wildner, W.; Güthner, T.; Holzrichter, K.; Mertschenk, B.; Ulrich, A. Safe Transport of Cyanamide. J. Hazard. Mater. 2009, 170, 829–835. [Google Scholar] [CrossRef] [PubMed]
  57. Pham, T.-T.; Shin, E.W. Influence of g-C3N4 Precursors in g-C3N4/NiTiO3 Composites on Photocatalytic Behavior and the Interconnection between g-C3N4 and NiTiO3. Langmuir 2018, 34, 13144–13154. [Google Scholar] [CrossRef]
  58. Guan, C.; Jiang, J.; Pang, S.; Chen, X.; Webster, R.D.; Lim, T.-T. Facile Synthesis of Pure g-C3N4 Materials for Peroxymonosulfate Activation to Degrade Bisphenol A: Effects of Precursors and Annealing Ambience on Catalytic Oxidation. Chem. Eng. J. 2020, 387, 123726. [Google Scholar] [CrossRef]
  59. Jung, H.; Pham, T.-T.; Shin, E.W. Effect of g-C3N4 Precursors on the Morphological Structures of g-C3N4/ZnO Composite Photocatalysts. J. Alloys Compd. 2019, 788, 1084–1092. [Google Scholar] [CrossRef]
  60. Zhang, G.; Zhang, J.; Zhang, M.; Wang, X. Polycondensation of Thiourea into Carbon Nitride Semiconductors as Visible Light Photocatalysts. J. Mater. Chem. 2012, 22, 8083–8091. [Google Scholar] [CrossRef]
  61. Xue, J.; Ma, S.; Zhou, Y.; Wang, Q. Au-Loaded Porous Graphitic C3N4/Graphene Layered Composite as a Ternary Plasmonic Photocatalyst and Its Visible-Light Photocatalytic Performance. Rsc Adv. 2015, 5, 88249–88257. [Google Scholar] [CrossRef]
  62. Edgar, I.C.; Johnstone, S.M.; Milton, M. Urea to Melamine Process. U.S. Patent 3,095,416, 25 June 1963. [Google Scholar]
  63. Wen, J.; Xie, J.; Chen, X.; Li, X. A Review on g-C3N4 -Based Photocatalysts. Appl. Surf. Sci. 2017, 391, 72–123. [Google Scholar] [CrossRef]
  64. Mishra, A.; Mehta, A.; Basu, S.; Shetti, N.P.; Reddy, K.R.; Aminabhavi, T.M. Graphitic Carbon Nitride (g-C3N4)—Based Metal-Free Photocatalysts for Water Splitting: A Review. Carbon 2019, 149, 693–721. [Google Scholar] [CrossRef]
  65. Teixeira, I.F.; Barbosa, E.C.M.; Tsang, S.C.E.; Camargo, P.H.C. Carbon Nitrides and Metal Nanoparticles: From Controlled Synthesis to Design Principles for Improved Photocatalysis. Chem. Soc. Rev. 2018, 47, 7783–7817. [Google Scholar] [CrossRef] [PubMed]
  66. Kumar, S.; Karthikeyan, S.; Lee, A.F. g-C3N4-Based Nanomaterials for Visible Light-Driven Photocatalysis. Catalysts 2018, 8, 74. [Google Scholar] [CrossRef] [Green Version]
  67. Ai, Z.; Shao, Y.; Chang, B.; Zhang, L.; Shen, J.; Wu, Y.; Huang, B.; Hao, X. Rational Modulation of Pn Homojunction in P-Doped g-C3N4 Decorated with Ti3C2 for Photocatalytic Overall Water Splitting. Appl. Catal. B Environ. 2019, 259, 118077. [Google Scholar] [CrossRef]
  68. Pan, Y.; Li, D.; Jiang, H. Sodium-Doped C3N4/MOF Heterojunction Composites with Tunable Band Structures for Photocatalysis: Interplay between Light Harvesting and Electron Transfer. Chem. Eur. J. 2018, 24, 18403–18407. [Google Scholar] [CrossRef] [PubMed]
  69. Putri, L.K.; Ng, B.-J.; Er, C.-C.; Ong, W.-J.; Chang, W.S.; Mohamed, A.R.; Chai, S.-P. Insights on the Impact of Doping Levels in Oxygen-Doped GC3N4 and Its Effects on Photocatalytic Activity. Appl. Surf. Sci. 2020, 504, 144427. [Google Scholar] [CrossRef]
  70. Sudrajat, H. A One-Pot, Solid-State Route for Realizing Highly Visible Light Active Na-Doped GC3N4 Photocatalysts. J. Solid State Chem. 2018, 257, 26–33. [Google Scholar] [CrossRef]
  71. Wang, L.; Guo, X.; Chen, Y.; Ai, S.; Ding, H. Cobalt-Doped g-C3N4 as a Heterogeneous Catalyst for Photo-Assisted Activation of Peroxymonosulfate for the Degradation of Organic Contaminants. Appl. Surf. Sci. 2019, 467, 954–962. [Google Scholar] [CrossRef]
  72. Li, G.; Wang, R.; Wang, B.; Zhang, J. Sm-Doped Mesoporous g-C3N4 as Efficient Catalyst for Degradation of Tylosin: Influencing Factors and Toxicity Assessment. Appl. Surf. Sci. 2020, 517, 146212. [Google Scholar] [CrossRef]
  73. Bellardita, M.; García-López, E.I.; Marcì, G.; Krivtsov, I.; García, J.R.; Palmisano, L. Selective Photocatalytic Oxidation of Aromatic Alcohols in Water by Using P-Doped g-C3N4. Appl. Catal. B Environ. 2018, 220, 222–233. [Google Scholar] [CrossRef] [Green Version]
  74. Wu, M.; Zhang, J.; He, B.; Wang, H.; Wang, R.; Gong, Y. In-Situ Construction of Coral-like Porous P-Doped g-C3N4 Tubes with Hybrid 1D/2D Architecture and High Efficient Photocatalytic Hydrogen Evolution. Appl. Catal. B Environ. 2019, 241, 159–166. [Google Scholar] [CrossRef]
  75. Ranjbakhsh, E.; Izadyar, M.; Nakhaeipour, A.; Habibi-Yangjeh, A. P-doped g-C3N4 as an Efficient Photocatalyst for CO2 Conversion into Value-added Materials: A Joint Experimental and Theoretical Study. Int. J. Quantum Chem. 2020, 120, e26388. [Google Scholar] [CrossRef]
  76. Chen, P.; Chen, L.; Ge, S.; Zhang, W.; Wu, M.; Xing, P.; Rotamond, T.B.; Lin, H.; Wu, Y.; He, Y. Microwave Heating Preparation of Phosphorus Doped g-C3N4 and Its Enhanced Performance for Photocatalytic H2 Evolution in the Help of Ag3PO4 Nanoparticles. Int. J. Hydrogen Energy 2020, 45, 14354–14367. [Google Scholar] [CrossRef]
  77. Cao, S.; Fan, B.; Feng, Y.; Chen, H.; Jiang, F.; Wang, X. Sulfur-Doped g-C3N4 Nanosheets with Carbon Vacancies: General Synthesis and Improved Activity for Simulated Solar-Light Photocatalytic Nitrogen Fixation. Chem. Eng. J. 2018, 353, 147–156. [Google Scholar] [CrossRef]
  78. Chen, Y.; Li, W.; Jiang, D.; Men, K.; Li, Z.; Li, L.; Sun, S.; Li, J.; Huang, Z.-H.; Wang, L.-N. Facile Synthesis of Bimodal Macroporous g-C3N4/SnO2 Nanohybrids with Enhanced Photocatalytic Activity. Sci. Bull. 2019, 64, 44–53. [Google Scholar] [CrossRef] [Green Version]
  79. Long, D.; Chen, Z.; Rao, X.; Zhang, Y. Sulfur-Doped g-C3N4 and BiPO4 Nanorod Hybrid Architectures for Enhanced Photocatalytic Hydrogen Evolution under Visible Light Irradiation. ACS Appl. Energy Mater. 2020, 3, 5024–5030. [Google Scholar]
  80. Oh, W.-D.; Lok, L.-W.; Veksha, A.; Giannis, A.; Lim, T.-T. Enhanced Photocatalytic Degradation of Bisphenol A with Ag-Decorated S-Doped g-C3N4 under Solar Irradiation: Performance and Mechanistic Studies. Chem. Eng. J. 2018, 333, 739–749. [Google Scholar] [CrossRef]
  81. Sun, C.; Zhang, H.; Liu, H.; Zheng, X.; Zou, W.; Dong, L.; Qi, L. Enhanced Activity of Visible-Light Photocatalytic H2 Evolution of Sulfur-Doped g-C3N4 Photocatalyst via Nanoparticle Metal Ni as Cocatalyst. Appl. Catal. B Environ. 2018, 235, 66–74. [Google Scholar] [CrossRef]
  82. Fu, J.; Zhu, B.; Jiang, C.; Cheng, B.; You, W.; Yu, J. Hierarchical Porous O-doped g-C3N4 with Enhanced Photocatalytic CO2 Reduction Activity. Small 2017, 13, 1603938. [Google Scholar] [CrossRef] [PubMed]
  83. Wang, H.; Guan, Y.; Hu, S.; Pei, Y.; Ma, W.; Fan, Z. Hydrothermal Synthesis of Band Gap-Tunable Oxygen-Doped g-C3N4 with Outstanding “Two-Channel” Photocatalytic H2O2 Production Ability Assisted by Dissolution–Precipitation Process. Nano 2019, 14, 1950023. [Google Scholar] [CrossRef]
  84. Zeng, Y.; Liu, X.; Liu, C.; Wang, L.; Xia, Y.; Zhang, S.; Luo, S.; Pei, Y. Scalable One-Step Production of Porous Oxygen-Doped g-C3N4 Nanorods with Effective Electron Separation for Excellent Visible-Light Photocatalytic Activity. Appl. Catal. B Environ. 2018, 224, 1–9. [Google Scholar] [CrossRef]
  85. Zhang, Y.; Chen, Z.; Li, J.; Lu, Z.; Wang, X. Self-Assembled Synthesis of Oxygen-Doped g-C3N4 Nanotubes in Enhancement of Visible-Light Photocatalytic Hydrogen. J. Energy Chem. 2021, 54, 36–44. [Google Scholar] [CrossRef]
  86. Panneri, S.; Ganguly, P.; Mohan, M.; Nair, B.N.; Mohamed, A.A.P.; Warrier, K.G.; Hareesh, U.S. Photoregenerable, Bifunctional Granules of Carbon-Doped g-C3N4 as Adsorptive Photocatalyst for the Efficient Removal of Tetracycline Antibiotic. ACS Sustain. Chem. Eng. 2017, 5, 1610–1618. [Google Scholar] [CrossRef]
  87. Babu, P.; Mohanty, S.; Naik, B.; Parida, K. Serendipitous Assembly of Mixed Phase BiVO4 on B-Doped g-C3N4: An Appropriate p–n Heterojunction for Photocatalytic O2 Evolution and Cr (VI) Reduction. Inorg. Chem. 2019, 58, 12480–12491. [Google Scholar] [CrossRef]
  88. Jiang, L.; Yuan, X.; Zeng, G.; Liang, J.; Wu, Z.; Yu, H.; Mo, D.; Wang, H.; Xiao, Z.; Zhou, C. Nitrogen Self-Doped g-C3N4 Nanosheets with Tunable Band Structures for Enhanced Photocatalytic Tetracycline Degradation. J. Colloid Interface Sci. 2019, 536, 17–29. [Google Scholar] [CrossRef] [PubMed]
  89. Zhu, D.; Zhou, Q. Nitrogen Doped g-C3N4 with the Extremely Narrow Band Gap for Excellent Photocatalytic Activities under Visible Light. Appl. Catal. B Environ. 2021, 281, 119474. [Google Scholar] [CrossRef]
  90. Wang, S.; Zhan, J.; Chen, K.; Ali, A.; Zeng, L.; Zhao, H.; Hu, W.; Zhu, L.; Xu, X. Potassium-Doped g-C3N4 Achieving Efficient Visible-Light-Driven CO2 Reduction. ACS Sustain. Chem. Eng. 2020, 8, 8214–8222. [Google Scholar] [CrossRef]
  91. Zhang, R.; Niu, S.; Zhang, X.; Jiang, Z.; Zheng, J.; Guo, C. Combination of Experimental and Theoretical Investigation on Ti-Doped g-C3N4 with Improved Photo-Catalytic Activity. Appl. Surf. Sci. 2019, 489, 427–434. [Google Scholar] [CrossRef]
  92. Wang, J.-C.; Cui, C.-X.; Kong, Q.-Q.; Ren, C.-Y.; Li, Z.; Qu, L.; Zhang, Y.; Jiang, K. Mn-Doped g-C3N4 Nanoribbon for Efficient Visible-Light Photocatalytic Water Splitting Coupling with Methylene Blue Degradation. ACS Sustain. Chem. Eng. 2018, 6, 8754–8761. [Google Scholar] [CrossRef]
  93. Faisal, M.; Ismail, A.A.; Harraz, F.A.; Al-Sayari, S.A.; El-Toni, A.M.; Al-Assiri, M.S. Synthesis of Highly Dispersed Silver Doped g-C3N4 Nanocomposites with Enhanced Visible-Light Photocatalytic Activity. Mater. Des. 2016, 98, 223–230. [Google Scholar] [CrossRef]
  94. Hu, J.; Zhang, P.; An, W.; Liu, L.; Liang, Y.; Cui, W. In-Situ Fe-Doped g-C3N4 Heterogeneous Catalyst via Photocatalysis-Fenton Reaction with Enriched Photocatalytic Performance for Removal of Complex Wastewater. Appl. Catal. B Environ. 2019, 245, 130–142. [Google Scholar] [CrossRef]
  95. Babu, B.; Shim, J.; Yoo, K. Efficient Solar-Light-Driven Photoelectrochemical Water Oxidation of One-Step in-Situ Synthesized Co-Doped g-C3N4 Nanolayers. Ceram. Int. 2020, 46, 16422–16430. [Google Scholar] [CrossRef]
  96. Huang, J.; Li, D.; Li, R.; Zhang, Q.; Chen, T.; Liu, H.; Liu, Y.; Lv, W.; Liu, G. An Efficient Metal-Free Phosphorus and Oxygen Co-Doped g-C3N4 Photocatalyst with Enhanced Visible Light Photocatalytic Activity for the Degradation of Fluoroquinolone Antibiotics. Chem. Eng. J. 2019, 374, 242–253. [Google Scholar] [CrossRef]
  97. Zhou, P.; Meng, X.; Li, L.; Sun, T. P, S Co-Doped g-C3N4 Isotype Heterojunction Composites for High-Efficiency Photocatalytic H2 Evolution. J. Alloys Compd. 2020, 827, 154259. [Google Scholar] [CrossRef]
  98. Cui, Y.; Wang, H.; Yang, C.; Li, M.; Zhao, Y.; Chen, F. Post-Activation of in Situ BF Codoped g-C3N4 for Enhanced Photocatalytic H2 Evolution. Appl. Surf. Sci. 2018, 441, 621–630. [Google Scholar] [CrossRef]
  99. Fang, W.; Liu, J.; Yu, L.; Jiang, Z.; Shangguan, W. Novel (Na, O) Co-Doped g-C3N4 with Simultaneously Enhanced Absorption and Narrowed Bandgap for Highly Efficient Hydrogen Evolution. Appl. Catal. B Environ. 2017, 209, 631–636. [Google Scholar] [CrossRef]
  100. Chen, C.; Xie, M.; Kong, L.; Lu, W.; Feng, Z.; Zhan, J. Mn3O4 Nanodots Loaded g-C3N4 Nanosheets for Catalytic Membrane Degradation of Organic Contaminants. J. Hazard. Mater. 2020, 390, 122146. [Google Scholar] [CrossRef] [PubMed]
  101. Surikanti, G.R.; Bajaj, P.; Sunkara, M.V. g-C3N4-Mediated Synthesis of Cu2O To Obtain Porous Composites with Improved Visible Light Photocatalytic Degradation of Organic Dyes. ACS Omega 2019, 4, 17301–17316. [Google Scholar] [CrossRef] [Green Version]
  102. Yan, D.; Wu, X.; Pei, J.; Wu, C.; Wang, X.; Zhao, H. Construction of g-C3N4/TiO2/Ag Composites with Enhanced Visible-Light Photocatalytic Activity and Antibacterial Properties. Ceram. Ceram. Int. 2020, 46, 696–702. [Google Scholar] [CrossRef]
  103. Rabani, I.; Zafar, R.; Subalakshmi, K.; Kim, H.-S.; Bathula, C.; Seo, Y.-S. A Facile Mechanochemical Preparation of Co3O4@ g-C3N4 for Application in Supercapacitors and Degradation of Pollutants in Water. J. Hazard. Mater. 2021, 407, 124360. [Google Scholar] [CrossRef] [PubMed]
  104. Mohammadi, R.; Alamgholiloo, H.; Gholipour, B.; Rostamnia, S.; Khaksar, S.; Farajzadeh, M.; Shokouhimehr, M. Visible-Light-Driven Photocatalytic Activity of ZnO/g-C3N4 Heterojunction for the Green Synthesis of Biologically Interest Small Molecules of Thiazolidinones. J. Photochem. Photobiol. A Chem. 2020, 402, 112786. [Google Scholar] [CrossRef]
  105. Bard, A.J.; Fox, M.A. Artificial Photosynthesis: Solar Splitting of Water to Hydrogen and Oxygen. Acc. Chem. Res. 1995, 28, 141–145. [Google Scholar] [CrossRef]
  106. Thompson, T.L.; Yates, J.T. Surface Science Studies of the Photoactivation of TiO2 New Photochemical Processes. Chem. Rev. 2006, 106, 4428–4453. [Google Scholar] [CrossRef] [PubMed]
  107. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  108. Tan, Y.; Chen, Y.; Mahimwalla, Z.; Johnson, M.B.; Sharma, T.; Brüning, R.; Ghandi, K. Novel Synthesis of Rutile Titanium Dioxide–Polypyrrole Nano Composites and Their Application in Hydrogen Generation. Synth. Met. 2014, 189, 77–85. [Google Scholar] [CrossRef]
  109. Christoforidis, K.C.; Montini, T.; Fittipaldi, M.; Jaén, J.J.D.; Fornasiero, P. Photocatalytic Hydrogen Production by Boron Modified TiO2/Carbon Nitride Heterojunctions. ChemCatChem 2019, 11, 6408–6416. [Google Scholar] [CrossRef]
  110. Crake, A.; Christoforidis, K.C.; Godin, R.; Moss, B.; Kafizas, A.; Zafeiratos, S.; Durrant, J.R.; Petit, C. Titanium Dioxide/Carbon Nitride Nanosheet Nanocomposites for Gas Phase CO2 Photoreduction under UV-Visible Irradiation. Appl. Catal. B Environ. 2019, 242, 369–378. [Google Scholar] [CrossRef]
  111. Acharya, R.; Parida, K. A Review on TiO2/g-C3N4 Visible-Light-Responsive Photocatalysts for Sustainable Energy Generation and Environmental Remediation. J. Environ. Chem. Eng. 2020, 8, 103896. [Google Scholar] [CrossRef]
  112. Alcudia-Ramos, M.A.; Fuentez-Torres, M.O.; Ortiz-Chi, F.; Espinosa-González, C.G.; Hernández-Como, N.; García-Zaleta, D.S.; Kesarla, M.K.; Torres-Torres, J.G.; Collins-Martínez, V.; Godavarthi, S. Fabrication of g-C3N4/TiO2 Heterojunction Composite for Enhanced Photocatalytic Hydrogen Production. Ceram. Int. 2020, 46, 38–45. [Google Scholar] [CrossRef]
  113. Li, G.; Nie, X.; Gao, Y.; An, T. Can Environmental Pharmaceuticals Be Photocatalytically Degraded and Completely Mineralized in Water Using g-C3N4/TiO2 under Visible Light Irradiation?—Implications of Persistent Toxic Intermediates. Appl. Catal. B Environ. 2016, 180, 726–732. [Google Scholar] [CrossRef]
  114. Li, G.; Nie, X.; Chen, J.; Jiang, Q.; An, T.; Wong, P.K.; Zhang, H.; Zhao, H.; Yamashita, H. Enhanced Visible-Light-Driven Photocatalytic Inactivation of Escherichia Coli Using g-C3N4/TiO2 Hybrid Photocatalyst Synthesized Using a Hydrothermal-Calcination Approach. Water Res. 2015, 86, 17–24. [Google Scholar] [CrossRef] [PubMed]
  115. Elbanna, O.; Fujitsuka, M.; Majima, T. g-C3N4/TiO2 Mesocrystals Composite for H2 Evolution under Visible-Light Irradiation and Its Charge Carrier Dynamics. ACS Appl. Mater. Interfaces 2017, 9, 34844–34854. [Google Scholar] [CrossRef] [PubMed]
  116. Kočí, K.; Reli, M.; Troppová, I.; Šihor, M.; Kupková, J.; Kustrowski, P.; Praus, P. Photocatalytic Decomposition of N2O over TiO2/g-C3N4 Photocatalysts Heterojunction. Appl. Surf. Sci. 2017, 396, 1685–1695. [Google Scholar] [CrossRef]
  117. Jiang, G.; Yang, X.; Wu, Y.; Li, Z.; Han, Y.; Shen, X. A Study of Spherical TiO2/g-C3N4 Photocatalyst: Morphology, Chemical Composition and Photocatalytic Performance in Visible Light. Mol. Catal. 2017, 432, 232–241. [Google Scholar] [CrossRef]
  118. Miranda, C.; Mansilla, H.; Yáñez, J.; Obregón, S.; Colón, G. Improved Photocatalytic Activity of g-C3N4/TiO2 Composites Prepared by a Simple Impregnation Method. J. Photochem. Photobiol. A Chem. 2013, 253, 16–21. [Google Scholar] [CrossRef]
  119. Zhang, L.; He, X.; Xu, X.; Liu, C.; Duan, Y.; Hou, L.; Zhou, Q.; Ma, C.; Yang, X.; Liu, R. Highly Active TiO2/g-C3N4/G Photocatalyst with Extended Spectral Response towards Selective Reduction of Nitrobenzene. Appl. Catal. B Environ. 2017, 203, 1–8. [Google Scholar] [CrossRef]
  120. Wang, J.; Wang, G.; Wang, X.; Wu, Y.; Su, Y.; Tang, H. 3D/2D Direct Z-Scheme Heterojunctions of Hierarchical TiO2 Microflowers/g-C3N4 Nanosheets with Enhanced Charge Carrier Separation for Photocatalytic H2 Evolution. Carbon 2019, 149, 618–626. [Google Scholar] [CrossRef]
  121. Boonprakob, N.; Wetchakun, N.; Phanichphant, S.; Waxler, D.; Sherrell, P.; Nattestad, A.; Chen, J.; Inceesungvorn, B. Enhanced Visible-Light Photocatalytic Activity of g-C3N4/TiO2 Films. J. Colloid Interface Sci. 2014, 417, 402–409. [Google Scholar] [CrossRef]
  122. Li, C.; Lou, Z.; Yang, Y.; Wang, Y.; Lu, Y.; Ye, Z.; Zhu, L. Hollowsphere Nanoheterojunction of g-C3N4@ TiO2 with High Visible Light Photocatalytic Property. Langmuir 2019, 35, 779–786. [Google Scholar] [CrossRef]
  123. Wei, H.; McMaster, W.A.; Tan, J.Z.Y.; Cao, L.; Chen, D.; Caruso, R.A. Mesoporous TiO2/g-C3N4 Microspheres with Enhanced Visible-Light Photocatalytic Activity. J. Phys. Chem. C 2017, 121, 22114–22122. [Google Scholar] [CrossRef]
  124. Zou, Y.; Yang, B.; Liu, Y.; Ren, Y.; Ma, J.; Zhou, X.; Cheng, X.; Deng, Y. Controllable Interface-Induced Co-Assembly toward Highly Ordered Mesoporous Pt@ TiO2/g-C3N4 Heterojunctions with Enhanced Photocatalytic Performance. Adv. Funct. Mater. 2018, 28, 1806214. [Google Scholar] [CrossRef]
  125. Li, K.; Huang, Z.; Zeng, X.; Huang, B.; Gao, S.; Lu, J. Synergetic Effect of Ti3+ and Oxygen Doping on Enhancing Photoelectrochemical and Photocatalytic Properties of TiO2/g-C3N4 Heterojunctions. ACS Appl. Mater. Interfaces 2017, 9, 11577–11586. [Google Scholar] [CrossRef] [PubMed]
  126. Rathi, A.K.; Kmentová, H.; Naldoni, A.; Goswami, A.; Gawande, M.B.; Varma, R.S.; Kment, S.; Zbořil, R. Significant Enhancement of Photoactivity in Hybrid TiO2/g-C3N4 Nanorod Catalysts Modified with Cu–Ni-Based Nanostructures. ACS Appl. Nano Mater. 2018, 1, 2526–2535. [Google Scholar] [CrossRef]
  127. Liu, C.; Raziq, F.; Li, Z.; Qu, Y.; Zada, A.; Jing, L. Synthesis of TiO2/g-C3N4 Nanocomposites with Phosphate–Oxygen Functional Bridges for Improved Photocatalytic Activity. Chin. J. Catal. 2017, 38, 1072–1078. [Google Scholar] [CrossRef]
  128. Li, Y.; Wang, R.; Li, H.; Wei, X.; Feng, J.; Liu, K.; Dang, Y.; Zhou, A. Efficient and Stable Photoelectrochemical Seawater Splitting with TiO2@g-C3N4 Nanorod Arrays Decorated by Co-Pi. J. Phys. Chem. C 2015, 119, 20283–20292. [Google Scholar] [CrossRef]
  129. Kholikov, B.; Hussain, J.; Zeng, H. Gold Modified TiO2/g-C3N4 for Enhanced Photocatalytic Activities to Evolved H2 Fuel. Inorg. Chem. Commun. 2021, 131, 108787. [Google Scholar] [CrossRef]
  130. Wang, C.; Zhao, Y.; Xu, H.; Li, Y.; Wei, Y.; Liu, J.; Zhao, Z. Efficient Z-Scheme Photocatalysts of Ultrathin g-C3N4-Wrapped Au/TiO2-Nanocrystals for Enhanced Visible-Light-Driven Conversion of CO2 with H2O. Appl. Catal. B Environ. 2020, 263, 118314. [Google Scholar] [CrossRef]
  131. Zada, A.; Qu, Y.; Ali, S.; Sun, N.; Lu, H.; Yan, R.; Zhang, X.; Jing, L. Improved Visible-Light Activities for Degrading Pollutants on TiO2/g-C3N4 Nanocomposites by Decorating SPR Au Nanoparticles and 2, 4-Dichlorophenol Decomposition Path. J. Hazard. Mater. 2018, 342, 715–723. [Google Scholar] [CrossRef] [PubMed]
  132. Wang, C.; Rao, Z.; Mahmood, A.; Wang, X.; Wang, Y.; Xie, X.; Sun, J. Improved Photocatalytic Oxidation Performance of Gaseous Acetaldehyde by Ternary g-C3N4/Ag-TiO2 Composites under Visible Light. J. Colloid Interface Sci. 2021, 602, 699–711. [Google Scholar] [CrossRef]
  133. Wang, W.; Zhang, D.; Sun, P.; Ji, Z.; Duan, J. High Efficiency Photocatalytic Degradation of Indoor Formaldehyde by Ag/g-C3N4/TiO2 Composite Catalyst with ZSM-5 as the Carrier. Microporous Mesoporous Mater. 2021, 322, 111134. [Google Scholar] [CrossRef]
  134. Yu, B.; Meng, F.; Khan, M.W.; Qin, R.; Liu, X. Facile Synthesis of AgNPs Modified TiO2@ g-C3N4 Heterojunction Composites with Enhanced Photocatalytic Activity under Simulated Sunlight. Mater. Res. Bull. 2020, 121, 110641. [Google Scholar] [CrossRef]
  135. Chen, Y.; Huang, W.; He, D.; Situ, Y.; Huang, H. Construction of Heterostructured g-C3N4/Ag/TiO2 Microspheres with Enhanced Photocatalysis Performance under Visible-Light Irradiation. ACS Appl. Mater. Interfaces 2014, 6, 14405–14414. [Google Scholar] [CrossRef] [PubMed]
  136. Li, G.; Wu, Y.; Zhang, M.; Chu, B.; Huang, W.; Fan, M.; Dong, L.; Li, B. Enhanced Removal of Toxic Cr (VI) in Wastewater by Synthetic TiO2/g-C3N4 Microspheres/RGO Photocatalyst under Irradiation of Visible Light. Ind. Eng. Chem. Res. 2019, 58, 8979–8989. [Google Scholar] [CrossRef]
  137. Zhang, C.; Zhou, Y.; Bao, J.; Fang, J.; Zhao, S.; Zhang, Y.; Sheng, X.; Chen, W. Structure Regulation of ZnS@ g-C3N4/TiO2 Nanospheres for Efficient Photocatalytic H2 Production under Visible-Light Irradiation. Chem. Eng. J. 2018, 346, 226–237. [Google Scholar] [CrossRef]
  138. Hu, M.; Xing, Z.; Cao, Y.; Li, Z.; Yan, X.; Xiu, Z.; Zhao, T.; Yang, S.; Zhou, W. Ti3+ Self-Doped Mesoporous Black TiO2/SiO2/g-C3N4 Sheets Heterojunctions as Remarkable Visible-Lightdriven Photocatalysts. Appl. Catal. B Environ. 2018, 226, 499–508. [Google Scholar] [CrossRef]
  139. Obregón, S.; Zhang, Y.; Colón, G. Cascade Charge Separation Mechanism by Ternary Heterostructured BiPO4/TiO2/g-C3N4 Photocatalyst. Appl. Catal. B Environ. 2016, 184, 96–103. [Google Scholar] [CrossRef]
  140. Zhao, H.; Tian, C.; Mei, J.; Yang, S.; Wong, P.K. Faster Electron Injection and Higher Interface Reactivity in g-C3N4/Fe2O3 Nanohybrid for Efficient Photo-Fenton-like Activity toward Antibiotics Degradation. Environ. Res. 2021, 195, 110842. [Google Scholar] [CrossRef] [PubMed]
  141. Sun, X.-Y.; Zhang, F.-J.; Kong, C. Porous g-C3N4/WO3 Photocatalyst Prepared by Simple Calcination for Efficient Hydrogen Generation under Visible Light. Colloids Surfaces A Physicochem. Eng. Asp. 2020, 594, 124653. [Google Scholar] [CrossRef]
  142. Zhang, Z.; Sun, Y.; Wang, Y.; Yang, Y.; Wang, P.; Shi, L.; Feng, L.; Fang, S.; Liu, Q.; Ma, L. Synthesis and Photocatalytic Activity of g-C3N4/ZnO Composite Microspheres under Visible Light Exposure. Ceram. Int. 2022, 48, 3293–3302. [Google Scholar] [CrossRef]
  143. Moradi, S.; Isari, A.A.; Hayati, F.; Kalantary, R.R.; Kakavandi, B. Co-Implanting of TiO2 and Liquid-Phase-Delaminated g-C3N4 on Multi-Functional Graphene Nanobridges for Enhancing Photocatalytic Degradation of Acetaminophen. Chem. Eng. J. 2021, 414, 128618. [Google Scholar] [CrossRef]
  144. Jo, W.-K.; Natarajan, T.S. Influence of TiO2 Morphology on the Photocatalytic Efficiency of Direct Z-Scheme g-C3N4/TiO2 Photocatalysts for Isoniazid Degradation. Chem. Eng. J. 2015, 281, 549–565. [Google Scholar] [CrossRef]
  145. Saravanan, R.; Gupta, V.K.; Mosquera, E.; Gracia, F. Preparation and Characterization of V2O5/ZnO Nanocomposite System for Photocatalytic Application. J. Mol. Liq. 2014, 198, 409–412. [Google Scholar] [CrossRef]
  146. Landry, C.J.; Burns, F.P.; Baerlocher, F.; Ghandi, K. Novel Solid-State Microbial Sensors Based on ZnO Nanorod Arrays. Adv. Funct. Mater. 2018, 28, 1706309. [Google Scholar] [CrossRef]
  147. Kumar, S.; Baruah, A.; Tonda, S.; Kumar, B.; Shanker, V.; Sreedhar, B. Cost-Effective and Eco-Friendly Synthesis of Novel and Stable N-Doped ZnO/g-C3N4 Core–Shell Nanoplates with Excellent Visible-Light Responsive Photocatalysis. Nanoscale 2014, 6, 4830–4842. [Google Scholar] [CrossRef] [Green Version]
  148. Uma, R.; Ravichandran, K.; Sriram, S.; Sakthivel, B. Cost-Effective Fabrication of ZnO/g-C3N4 Composite Thin Films for Enhanced Photocatalytic Activity against Three Different Dyes (MB, MG and RhB). Mater. Chem. Phys. 2017, 201, 147–155. [Google Scholar] [CrossRef]
  149. Zhang, N.; Gao, J.; Huang, C.; Liu, W.; Tong, P.; Zhang, L. In Situ Hydrothermal Growth of ZnO/g-C3N4 Nanoflowers Coated Solid-Phase Microextraction Fibers Coupled with GC-MS for Determination of Pesticides Residues. Anal. Chim. Acta 2016, 934, 122–131. [Google Scholar] [CrossRef]
  150. Mohammadi, I.; Zeraatpisheh, F.; Ashiri, E.; Abdi, K. Solvothermal Synthesis of g-C3N4 and ZnO Nanoparticles on TiO2 Nanotube as Photoanode in DSSC. Int. J. Hydrogen Energy 2020, 45, 18831–18839. [Google Scholar] [CrossRef]
  151. Jang, E.; Kim, D.W.; Hong, S.H.; Park, Y.M.; Park, T.J. Visible Light-Driven g-C3N4@ ZnO Heterojunction Photocatalyst Synthesized via Atomic Layer Deposition with a Specially Designed Rotary Reactor. Appl. Surf. Sci. 2019, 487, 206–210. [Google Scholar] [CrossRef]
  152. Fageria, P.; Nazir, R.; Gangopadhyay, S.; Barshilia, H.C.; Pande, S. Graphitic-Carbon Nitride Support for the Synthesis of Shape-Dependent ZnO and Their Application in Visible Light Photocatalysts. Rsc Adv. 2015, 5, 80397–80409. [Google Scholar] [CrossRef]
  153. Tan, X.; Wang, X.; Hang, H.; Zhang, D.; Zhang, N.; Xiao, Z.; Tao, H. Self-Assembly Method Assisted Synthesis of g-C3N4/ZnO Heterostructure Nanocomposites with Enhanced Photocatalytic Performance. Opt. Mater. 2019, 96, 109266. [Google Scholar] [CrossRef]
  154. Jung, H.; Pham, T.-T.; Shin, E.W. Interactions between ZnO Nanoparticles and Amorphous g-C3N4 Nanosheets in Thermal Formation of g-C3N4/ZnO Composite Materials: The Annealing Temperature Effect. Appl. Surf. Sci. 2018, 458, 369–381. [Google Scholar] [CrossRef]
  155. Liu, W.; Wang, M.; Xu, C.; Chen, S. Facile Synthesis of g-C3N4/ZnO Composite with Enhanced Visible Light Photooxidation and Photoreduction Properties. Chem. Eng. J. 2012, 209, 386–393. [Google Scholar] [CrossRef]
  156. Zhang, X.; Ling, S.; Ji, H.; Xu, L.; Huang, Y.; Hua, M.; Xia, J.; Li, H. Metal Ion-Containing Ionic Liquid Assisted Synthesis and Enhanced Photoelectrochemical Performance of g-C3N4/ZnO Composites. Mater. Technol. 2018, 33, 185–192. [Google Scholar] [CrossRef]
  157. Zhou, J.; Zhang, M.; Zhu, Y. Preparation of Visible Light-Driven g-C3N4@ ZnO Hybrid Photocatalyst via Mechanochemistry. Phys. Chem. Chem. Phys. 2014, 16, 17627–17633. [Google Scholar] [CrossRef] [PubMed]
  158. Paul, D.R.; Gautam, S.; Panchal, P.; Nehra, S.P.; Choudhary, P.; Sharma, A. ZnO-Modified g-C3N4: A Potential Photocatalyst for Environmental Application. ACS Omega 2020, 5, 3828–3838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Kuang, P.-Y.; Su, Y.-Z.; Chen, G.-F.; Luo, Z.; Xing, S.-Y.; Li, N.; Liu, Z.-Q. g-C3N4 Decorated ZnO Nanorod Arrays for Enhanced Photoelectrocatalytic Performance. Appl. Surf. Sci. 2015, 358, 296–303. [Google Scholar] [CrossRef]
  160. Ramachandra, M.; Devi Kalathiparambil Rajendra Pai, S.; Resnik Jaleel UC, J.; Pinheiro, D. Improved Photocatalytic Activity of g-C3N4/ZnO: A Potential Direct Z-Scheme Nanocomposite. ChemistrySelect 2020, 5, 11986–11995. [Google Scholar] [CrossRef]
  161. Xing, H.; Ma, H.; Fu, Y.; Xue, M.; Zhang, X.; Dong, X.; Zhang, X. Preparation of g-C3N4/ZnO Composites and Their Enhanced Photocatalytic Activity. Mater. Technol. 2015, 30, 122–127. [Google Scholar] [CrossRef]
  162. Ismael, M. The Photocatalytic Performance of the ZnO/g-C3N4 Composite Photocatalyst toward Degradation of Organic Pollutants and Its Inactivity toward Hydrogen Evolution: The Influence of Light Irradiation and Charge Transfer. Chem. Phys. Lett. 2020, 739, 136992. [Google Scholar] [CrossRef]
  163. Fang, Q.; Li, B.; Li, Y.-Y.; Huang, W.-Q.; Peng, W.; Fan, X.; Huang, G.-F. 0D/2D Z-Scheme Heterojunctions of g-C3N4 Quantum Dots/ZnO Nanosheets as a Highly Efficient Visible-Light Photocatalyst. Adv. Powder Technol. 2019, 30, 1576–1583. [Google Scholar] [CrossRef]
  164. Wang, J.; Yang, Z.; Gao, X.; Yao, W.; Wei, W.; Chen, X.; Zong, R.; Zhu, Y. Core-Shell g-C3N4@ ZnO Composites as Photoanodes with Double Synergistic Effects for Enhanced Visible-Light Photoelectrocatalytic Activities. Appl. Catal. B Environ. 2017, 217, 169–180. [Google Scholar] [CrossRef]
  165. Mathialagan, A.; Manavalan, M.; Venkatachalam, K.; Mohammad, F.; Oh, W.C.; Sagadevan, S. Fabrication and Physicochemical Characterization of g-C3N4/ZnO Composite with Enhanced Photocatalytic Activity under Visible Light. Opt. Mater. 2020, 100, 109643. [Google Scholar] [CrossRef]
  166. Liu, W.; Wang, M.; Xu, C.; Chen, S.; Fu, X. Significantly Enhanced Visible-Light Photocatalytic Activity of g-C3N4 via ZnO Modification and the Mechanism Study. J. Mol. Catal. A Chem. 2013, 368, 9–15. [Google Scholar] [CrossRef]
  167. Wang, L.; Ma, C.; Guo, Z.; Lv, Y.; Chen, W.; Chang, Z.; Yuan, Q.; Ming, H.; Wang, J. In-Situ Growth of g-C3N4 Layer on ZnO Nanoparticles with Enhanced Photocatalytic Performances under Visible Light Irradiation. Mater. Lett. 2017, 188, 347–350. [Google Scholar] [CrossRef] [Green Version]
  168. Zhong, Q.; Lan, H.; Zhang, M.; Zhu, H.; Bu, M. Preparation of Heterostructure g-C3N4/ZnO Nanorods for High Photocatalytic Activity on Different Pollutants (MB, RhB, Cr (VI) and Eosin). Ceram. Int. 2020, 46, 12192–12199. [Google Scholar] [CrossRef]
  169. Ahmad, I. Comparative Study of Metal (Al, Mg, Ni, Cu and Ag) Doped ZnO/g-C3N4 Composites: Efficient Photocatalysts for the Degradation of Organic Pollutants. Sep. Purif. Technol. 2020, 251, 117372. [Google Scholar] [CrossRef]
  170. Ta, Q.T.H.; Namgung, G.; Noh, J.-S. Facile Synthesis of Porous Metal-Doped ZnO/g-C3N4 Composites for Highly Efficient Photocatalysts. J. Photochem. Photobiol. A Chem. 2019, 368, 110–119. [Google Scholar]
  171. Jin, C.; Li, W.; Chen, Y.; Li, R.; Huo, J.; He, Q.; Wang, Y. Efficient Photocatalytic Degradation and Adsorption of Tetracycline over Type-II Heterojunctions Consisting of ZnO Nanorods and K-Doped Exfoliated g-C3N4 Nanosheets. Ind. Eng. Chem. Res. 2020, 59, 2860–2873. [Google Scholar] [CrossRef]
  172. Qamar, M.A.; Shahid, S.; Javed, M.; Iqbal, S.; Sher, M.; Akbar, M.B. Highly Efficient g-C3N4/Cr-ZnO Nanocomposites with Superior Photocatalytic and Antibacterial Activity. J. Photochem. Photobiol. A Chem. 2020, 401, 112776. [Google Scholar] [CrossRef]
  173. Neena, D.; Humayun, M.; Zuo, W.; Liu, C.S.; Gao, W.; Fu, D.J. Hierarchical Hetero-Architectures of in-Situ g-C3N4-Coupled Fe-Doped ZnO Micro-Flowers with Enhanced Visible-Light Photocatalytic Activities. Appl. Surf. Sci. 2020, 506, 145017. [Google Scholar] [CrossRef]
  174. Sun, M.; Chen, Z.; Jiang, X.; Feng, C.; Zeng, R. Optimized Preparation of Co-Pi Decorated g-C3N4 @ ZnO Shell-Core Nanorod Array for Its Improved Photoelectrochemical Performance and Stability. J. Alloys Compd. 2019, 780, 540–551. [Google Scholar] [CrossRef]
  175. Kumar, S.; Kumar, A.; Kumar, A.; Balaji, R.; Krishnan, V. Highly Efficient Visible Light Active 2D-2D Nanocomposites of N-ZnO-g-C3N4 for Photocatalytic Degradation of Diverse Industrial Pollutants. ChemistrySelect 2018, 3, 1919–1932. [Google Scholar] [CrossRef]
  176. Liu, X.; Liu, L.; Yao, Z.; Yang, Z.; Xu, H. Enhanced Visible-Light-Driven Photocatalytic Hydrogen Evolution and NO Photo-Oxidation Capacity of ZnO/g-C3N4 with N Dopant. Colloids Surfaces A Physicochem. Eng. Asp. 2020, 599, 124869. [Google Scholar] [CrossRef]
  177. Mohamed, M.A.; Zain, M.F.M.; Minggu, L.J.; Kassim, M.B.; Jaafar, J.; Amin, N.A.S.; Mastuli, M.S.; Wu, H.; Wong, R.J.; Ng, Y.H. Bio-Inspired Hierarchical Hetero-Architectures of in-Situ C-Doped g-C3N4 Grafted on C, N Co-Doped ZnO Micro-Flowers with Booming Solar Photocatalytic Activity. J. Ind. Eng. Chem. 2019, 77, 393–407. [Google Scholar] [CrossRef]
  178. Kong, J.-Z.; Zhai, H.-F.; Zhang, W.; Wang, S.-S.; Zhao, X.-R.; Li, M.; Li, H.; Li, A.-D.; Wu, D. Visible Light-Driven Photocatalytic Performance of N-Doped ZnO/g-C3N4 Nanocomposites. Nanoscale Res. Lett. 2017, 12, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Qin, J.; Yang, C.; Cao, M.; Zhang, X.; Rajendran, S.; Limpanart, S.; Ma, M.; Liu, R. Two-Dimensional Porous Sheet-like Carbon-Doped ZnO/g-C3N4 Nanocomposite with High Visible-Light Photocatalytic Performance. Mater. Lett. 2017, 189, 156–159. [Google Scholar] [CrossRef]
  180. Zhu, Y.-P.; Li, M.; Liu, Y.-L.; Ren, T.-Z.; Yuan, Z.-Y. Carbon-Doped ZnO Hybridized Homogeneously with Graphitic Carbon Nitride Nanocomposites for Photocatalysis. J. Phys. Chem. C 2014, 118, 10963–10971. [Google Scholar] [CrossRef]
  181. Xing, Z.; Chen, Y.; Liu, C.; Yang, J.; Xu, J.; Situ, Y.; Huang, H. Synthesis of Core-Shell ZnO/Oxygen Doped g-C3N4 Visible Light Driven Photocatalyst via Hydrothermal Method. J. Alloys Compd. 2017, 708, 853–861. [Google Scholar] [CrossRef]
  182. Chen, X.-Y.; Pan, Q.-J.; Guo, Y.-R. One-Pot Synthesis of the g-C3N4/S-Doped ZnO Composite with Assistance of Sodium Lignosulfonate and Its Enhanced Photodegradation on Organic Pollutants. Mater. Res. Bull. 2018, 107, 164–170. [Google Scholar] [CrossRef]
  183. Liu, S.; Li, F.; Li, Y.; Hao, Y.; Wang, X.; Li, B.; Liu, R. Fabrication of Ternary g-C3N4/Al2O3/ZnO Heterojunctions Based on Cascade Electron Transfer toward Molecular Oxygen Activation. Appl. Catal. B Environ. 2017, 212, 115–128. [Google Scholar] [CrossRef]
  184. Mohamed, N.A.; Safaei, J.; Ismail, A.F.; Khalid, M.N.; Jailani, M.F.A.M.; Noh, M.F.M.; Arzaee, N.A.; Zhou, D.; Sagu, J.S.; Teridi, M.A.M. Boosting Photocatalytic Activities of BiVO4 by Creation of g-C3N4/ZnO@ BiVO4 Heterojunction. Mater. Res. Bull. 2020, 125, 110779. [Google Scholar] [CrossRef]
  185. Patnaik, S.; Sahoo, D.P.; Mohapatra, L.; Martha, S.; Parida, K. ZnCr2O4@ ZnO/g-C3N4: A Triple-Junction Nanostructured Material for Effective Hydrogen and Oxygen Evolution under Visible Light. Energy Technol. 2017, 5, 1687–1701. [Google Scholar] [CrossRef]
  186. Dong, Z.; Wu, Y.; Thirugnanam, N.; Li, G. Double Z-Scheme ZnO/ZnS/g-C3N4 Ternary Structure for Efficient Photocatalytic H2 Production. Appl. Surf. Sci. 2018, 430, 293–300. [Google Scholar] [CrossRef]
  187. Xiao, J.; Zhang, X.; Li, Y. A Ternary g-C3N4/Pt/ZnO Photoanode for Efficient Photoelectrochemical Water Splitting. Int. J. Hydrogen Energy 2015, 40, 9080–9087. [Google Scholar] [CrossRef]
  188. Azimi, E.B.; Badiei, A.; Sadr, M.H. Dramatic Visible Photocatalytic Performance of g-C3N4 -Based Nanocomposite Due to the Synergistic Effect of AgBr and ZnO Semiconductors. J. Phys. Chem. Solids 2018, 122, 174–183. [Google Scholar] [CrossRef]
  189. Adhikari, S.P.; Pant, H.R.; Kim, J.H.; Kim, H.J.; Park, C.H.; Kim, C.S. One Pot Synthesis and Characterization of Ag-ZnO/g-C3N4 Photocatalyst with Improved Photoactivity and Antibacterial Properties. Colloids Surfaces A Physicochem. Eng. Asp. 2015, 482, 477–484. [Google Scholar] [CrossRef]
  190. Ma, S.; Zhan, S.; Xia, Y.; Wang, P.; Hou, Q.; Zhou, Q. Enhanced Photocatalytic Bactericidal Performance and Mechanism with Novel Ag/ZnO/g-C3N4 Composite under Visible Light. Catal. Today 2019, 330, 179–188. [Google Scholar] [CrossRef]
  191. Panneri, S.; Ganguly, P.; Nair, B.N.; Mohamed, A.A.P.; Warrier, K.G.; Hareesh, U.N.S. Copyrolysed C3N4-Ag/ZnO Ternary Heterostructure Systems for Enhanced Adsorption and Photocatalytic Degradation of Tetracycline. Eur. J. Inorg. Chem. 2016, 31, 5068–5076. [Google Scholar] [CrossRef]
  192. Lee, S.J.; Begildayeva, T.; Jung, H.J.; Koutavarapu, R.; Yu, Y.; Choi, M.; Choi, M.Y. Plasmonic ZnO/Au/g-C3N4 Nanocomposites as Solar Light Active Photocatalysts for Degradation of Organic Contaminants in Wastewater. Chemosphere 2021, 263, 128262. [Google Scholar] [CrossRef]
  193. Thang, N.Q.; Sabbah, A.; Chen, L.-C.; Chen, K.-H.; Thi, C.M.; Van Viet, P. Localized Surface Plasmonic Resonance Role of Silver Nanoparticles in the Enhancement of Long-Chain Hydrocarbons of the CO2 Reduction over Ag-GC3N4/ZnO Nanorods Photocatalysts. Chem. Eng. Sci. 2021, 229, 116049. [Google Scholar] [CrossRef]
  194. Zhang, Z.; Li, X.; Chen, H.; Shao, G.; Zhang, R.; Lu, H. Synthesis and Properties of Ag/ZnO/g-C3N4 Ternary Micro/Nano Composites by Microwave-Assisted Method. Mater. Res. Express 2018, 5, 15021. [Google Scholar] [CrossRef]
  195. Wei, X.; Liu, H.; Li, T.; Jiang, Z.; Hu, W.; Niu, Q.; Chen, J. Three-Dimensional Flower Heterojunction g-C3N4/Ag/ZnO Composed of Ultrathin Nanosheets with Enhanced Photocatalytic Performance. J. Photochem. Photobiol. A Chem. 2020, 390, 112342. [Google Scholar] [CrossRef]
  196. Rong, X.; Qiu, F.; Jiang, Z.; Rong, J.; Pan, J.; Zhang, T.; Yang, D. Preparation of Ternary Combined ZnO-Ag2O/Porous g-C3N4 Composite Photocatalyst and Enhanced Visible-Light Photocatalytic Activity for Degradation of Ciprofloxacin. Chem. Eng. Res. Des. 2016, 111, 253–261. [Google Scholar] [CrossRef]
  197. Zhu, P.; Hu, M.; Duan, M.; Xie, L.; Zhao, M. High Visible Light Response Z-Scheme Ag3PO4/g-C3N4/ZnO Composite Photocatalyst for Efficient Degradation of Tetracycline Hydrochloride: Preparation, Properties and Mechanism. J. Alloys Compd. 2020, 840, 155714. [Google Scholar] [CrossRef]
  198. Feng, Y.; Wang, Y.; Li, M.; Lv, S.; Li, W.; Li, Z. Novel Visible Light Induced Ag2S/g-C3N4/ZnO Nanoarrays Heterojunction for Efficient Photocatalytic Performance. Appl. Surf. Sci. 2018, 462, 896–903. [Google Scholar] [CrossRef]
  199. Kumaresan, N.; Sinthiya, M.M.A.; Kumar, M.P.; Ravichandran, S.; Babu, R.R.; Sethurman, K.; Ramamurthi, K. Investigation on the g-C3N4 Encapsulated ZnO Nanorods Heterojunction Coupled with GO for Effective Photocatalytic Activity under Visible Light Irradiation. Arab. J. Chem. 2020, 13, 2826–2843. [Google Scholar] [CrossRef]
  200. Iqbal, S.; Bahadur, A.; Javed, M.; Hakami, O.; Irfan, R.M.; Ahmad, Z.; AlObaid, A.; Al-Anazy, M.M.; Baghdadi, H.B.; Abd-Rabboh, H.S.M. Design Ag-Doped ZnO Heterostructure Photocatalyst with Sulfurized Graphitic C3N4 Showing Enhanced Photocatalytic Activity. Mater. Sci. Eng. B 2021, 272, 115320. [Google Scholar] [CrossRef]
  201. Xiang, Y.; Zhou, Q.; Li, Z.; Cui, Z.; Liu, X.; Liang, Y.; Zhu, S.; Zheng, Y.; Yeung, K.W.K.; Wu, S. A Z-Scheme Heterojunction of ZnO/CDots/C3N4 for Strengthened Photoresponsive Bacteria-Killing and Acceleration of Wound Healing. J. Mater. Sci. Technol. 2020, 57, 1–11. [Google Scholar] [CrossRef]
  202. Zhao, S.-W.; Zheng, M.; Sun, H.-L.; Li, S.-J.; Pan, Q.-J.; Guo, Y.-R. Construction of Heterostructured g-C3N4/ZnO/Cellulose and Its Antibacterial Activity: Experimental and Theoretical Investigations. Dalt. Trans. 2020, 49, 3723–3734. [Google Scholar] [CrossRef] [PubMed]
  203. Raza, W.; Faraz, M. Novel g-C3N4/Fe-ZnO/RGO Nanocomposites with Boosting Visible Light Photocatalytic Activity for MB, Cr (VI), and Outstanding Catalytic Activity toward Para-Nitrophenol Reduction. Nanotechnology 2020, 31, 325603. [Google Scholar] [CrossRef] [PubMed]
  204. Meng, F.; Chang, Y.; Qin, W.; Yuan, Z.; Zhao, J.; Zhang, J.; Han, E.; Wang, S.; Yang, M.; Shen, Y. ZnO-Reduced Graphene Oxide Composites Sensitized with Graphitic Carbon Nitride Nanosheets for Ethanol Sensing. ACS Appl. Nano Mater. 2019, 2, 2734–2742. [Google Scholar] [CrossRef]
  205. Guo, F.; Shi, W.; Guan, W.; Huang, H.; Liu, Y. Carbon Dots/g-C3N4/ZnO Nanocomposite as Efficient Visible-Light Driven Photocatalyst for Tetracycline Total Degradation. Sep. Purif. Technol. 2017, 173, 295–303. [Google Scholar] [CrossRef]
  206. Wang, J.; Luo, Z.; Song, Y.; Zheng, X.; Qu, L.; Qian, J.; Wu, Y.; Wu, X.; Wu, Z. Remediation of Phenanthrene Contaminated Soil by g-C3N4/Fe3O4 Composites and Its Phytotoxicity Evaluation. Chemosphere 2019, 221, 554–562. [Google Scholar] [CrossRef] [PubMed]
  207. Zhu, Z.; Yu, Y.; Dong, H.; Liu, Z.; Li, C.; Huo, P.; Yan, Y. Intercalation Effect of Attapulgite in g-C3N4 Modified with Fe3O4 Quantum Dots to Enhance Photocatalytic Activity for Removing 2-Mercaptobenzothiazole under Visible Light. ACS Sustain. Chem. Eng. 2017, 5, 10614–10623. [Google Scholar] [CrossRef]
  208. Lima, M.J.; Sampaio, M.J.; Silva, C.G.; Silva, A.M.T.; Faria, J.L. Magnetically Recoverable Fe3O4/g-C3N4 Composite for Photocatalytic Production of Benzaldehyde under UV-Led Radiation. Catal. Today 2019, 328, 293–299. [Google Scholar] [CrossRef]
  209. Wang, M.; Cui, S.; Yang, X.; Bi, W. Synthesis of g-C3N4/Fe3O4 Nanocomposites and Application as a New Sorbent for Solid Phase Extraction of Polycyclic Aromatic Hydrocarbons in Water Samples. Talanta 2015, 132, 922–928. [Google Scholar] [CrossRef]
  210. Xu, Q.; Zhu, B.; Jiang, C.; Cheng, B.; Yu, J. Constructing 2D/2D Fe2O3/g-C3N4 Direct Z-scheme Photocatalysts with Enhanced H2 Generation Performance. Sol. Rrl 2018, 2, 1800006. [Google Scholar] [CrossRef]
  211. Ye, S.; Qiu, L.-G.; Yuan, Y.-P.; Zhu, Y.-J.; Xia, J.; Zhu, J.-F. Facile Fabrication of Magnetically Separable Graphitic Carbon Nitride Photocatalysts with Enhanced Photocatalytic Activity under Visible Light. J. Mater. Chem. A 2013, 1, 3008–3015. [Google Scholar] [CrossRef]
  212. Kumar, S.; Jain, S.; Lamba, B.Y. A Review on Synthesis, Characterisation and Surface Modification of Magnetic Nanoparticle and Its Composite for Removal of Heavy Metals from Wastewater. Int. J. Environ. Anal. Chem. 2021, 1–16. [Google Scholar] [CrossRef]
  213. Ahmadi, R.; Hosseini, H.R.M. An Investigation on the Optimum Conditions of Synthesizing a Magnetite Based Ferrofluid as MRI Contrast Agent Using Taguchi Method. Mater. Sci. Pol. 2013, 31, 253–258. [Google Scholar] [CrossRef]
  214. Alaghmandfard, A.; Sedighi, O.; Rezaei, N.T.; Abedini, A.A.; Khachatourian, A.M.; Toprak, M.S.; Seifalian, A. Recent Advances in the Modification of Carbon-Based Quantum Dots for Biomedical Applications. Mater. Sci. Eng. C 2020, 120, 111756. [Google Scholar] [CrossRef]
  215. Alaghmandfard, A.; Madaah Hosseini, H.R. A Facile, Two-Step Synthesis and Characterization of Fe3O4–LCysteine–Graphene Quantum Dots as a Multifunctional Nanocomposite. Appl. Nanosci. 2021, 11, 849–860. [Google Scholar] [CrossRef]
  216. Liu, S.; Wang, S.; Jiang, Y.; Zhao, Z.; Jiang, G.; Sun, Z. Synthesis of Fe2O3 Loaded Porous g-C3N4 Photocatalyst for Photocatalytic Reduction of Dinitrogen to Ammonia. Chem. Eng. J. 2019, 373, 572–579. [Google Scholar] [CrossRef]
  217. Ding, Q.; Lam, F.L.Y.; Hu, X. Complete Degradation of Ciprofloxacin over g-C3N4 -Iron Oxide Composite via Heterogeneous Dark Fenton Reaction. J. Environ. Manag. 2019, 244, 23–32. [Google Scholar] [CrossRef] [PubMed]
  218. Papich, M.G. Ciprofloxacin Pharmacokinetics in Clinical Canine Patients. J. Vet. Intern. Med. 2017, 31, 1508–1513. [Google Scholar] [CrossRef]
  219. Duan, B.; Mei, L. A Z-Scheme Fe2O3/g-C3N4 Heterojunction for Carbon Dioxide to Hydrocarbon Fuel under Visible Illuminance. J. Colloid Interface Sci. 2020, 575, 265–273. [Google Scholar] [CrossRef] [PubMed]
  220. Geng, Y.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. Z-Scheme 2D/2D α-Fe2O3/g-C3N4 Heterojunction for Photocatalytic Oxidation of Nitric Oxide. Appl. Catal. B Environ. 2021, 280, 119409. [Google Scholar] [CrossRef]
  221. Cheng, R.; Zhang, L.; Fan, X.; Wang, M.; Li, M.; Shi, J. One-Step Construction of FeOx Modified g-C3N4 for Largely Enhanced Visible-Light Photocatalytic Hydrogen Evolution. Carbon 2016, 101, 62–70. [Google Scholar] [CrossRef]
  222. Wang, J.; Qin, C.; Wang, H.; Chu, M.; Zada, A.; Zhang, X.; Li, J.; Raziq, F.; Qu, Y.; Jing, L. Exceptional Photocatalytic Activities for CO2 Conversion on AlO Bridged g-C3N4/α-Fe2O3 z-Scheme Nanocomposites and Mechanism Insight with IsotopesZ. Appl. Catal. B Environ. 2018, 221, 459–466. [Google Scholar] [CrossRef]
  223. Christoforidis, K.C.; Montini, T.; Bontempi, E.; Zafeiratos, S.; Jaén, J.J.D.; Fornasiero, P. Synthesis and Photocatalytic Application of Visible-Light Active β-Fe2O3/g-C3N4 Hybrid Nanocomposites. Appl. Catal. B Environ. 2016, 187, 171–180. [Google Scholar] [CrossRef]
  224. Hassanzadeh-Tabrizi, S.A.; Nguyen, C.-C.; Do, T.-O. Synthesis of Fe2O3/Pt/Au Nanocomposite Immobilized on g-C3N4 for Localized Plasmon Photocatalytic Hydrogen Evolution. Appl. Surf. Sci. 2019, 489, 741–754. [Google Scholar] [CrossRef]
  225. Ghandi, K.; Wang, F.; Landry, C.; Mostafavi, M. Naked Gold Nanoparticles and Hot Electrons in Water. Sci. Rep. 2018, 8, 1–6. [Google Scholar] [CrossRef] [Green Version]
  226. Wei, Y.; Li, X.; Zhang, Y.; Yan, Y.; Huo, P.; Wang, H. g-C3N4 Quantum Dots and Au Nano Particles Co-Modified CeO2/Fe3O4 Micro-Flowers Photocatalyst for Enhanced CO2 Photoreduction. Renew. Energy 2021, 179, 756–765. [Google Scholar]
  227. Liu, L.; Wang, M.; Wang, C. In-Situ Synthesis of Graphitic Carbon Nitride/Iron Oxide−Copper Composites and Their Application in the Electrochemical Detection of Glucose. Electrochim. Acta 2018, 265, 275–283. [Google Scholar] [CrossRef]
  228. Zhang, M.; Huang, G.; Huang, J.; Chen, W. Three-Dimensional Multi-Walled Carbon Nanotubes@ g-C3N4@ Fe3O4nanocomposites-Based Magnetic Solid Phase Extraction for the Determination of Polycyclic Aromatic Hydrocarbons in Water Samples. Microchem. J. 2018, 142, 385–393. [Google Scholar] [CrossRef]
  229. Wang, L.; Sun, Q.; Dou, Y.; Zhang, Z.; Yan, T.; Li, Y. Fabricating a Novel Ternary Recyclable Fe3O4/Graphene/Sulfur-Doped g-C3N4 Composite Catalyst for Enhanced Removal of Ranitidine under Visible-Light Irradiation and Reducing of Its N-Nitrosodimethylamine Formation Potential. J. Hazard. Mater. 2021, 413, 125288. [Google Scholar] [CrossRef]
  230. Rashid, J.; Parveen, N.; Haq, T.U.; Iqbal, A.; Talib, S.H.; Awan, S.U.; Hussain, N.; Zaheer, M. g-C3N4/CeO2/Fe3O4Ternary Composite as an Efficient Bifunctional Catalyst for Overall Water Splitting. ChemCatChem 2018, 10, 5587–5592. [Google Scholar] [CrossRef]
  231. Zhu, Z.; Huo, P.; Lu, Z.; Yan, Y.; Liu, Z.; Shi, W.; Li, C.; Dong, H. Fabrication of Magnetically Recoverable Photocatalysts Using g-C3N4 for Effective Separation of Charge Carriers through like-Z-Scheme Mechanism with Fe3O4 Mediator. Chem. Eng. J. 2018, 331, 615–625. [Google Scholar] [CrossRef]
  232. Mishra, P.; Behera, A.; Kandi, D.; Ratha, S.; Parida, K. Novel Magnetic Retrievable Visible-Light-Driven Ternary Fe3O4@ NiFe2O4/Phosphorus-Doped g-C3N4 Nanocomposite Photocatalyst with Significantly Enhanced Activity through a Double-Z-Scheme System. Inorg. Chem. 2020, 59, 4255–4272. [Google Scholar] [CrossRef]
  233. Ji, H.; Jing, X.; Xu, Y.; Yan, J.; Li, H.; Li, Y.; Huang, L.; Zhang, Q.; Xu, H.; Li, H. Magnetic g-C3N4/NiFe2O4 Hybrids with Enhanced Photocatalytic Activity. RSC Adv. 2015, 5, 57960–57967. [Google Scholar] [CrossRef]
  234. Palanivel, B.; Ayappan, C.; Jayaraman, V.; Chidambaram, S.; Maheswaran, R.; Mani, A. Inverse Spinel NiFe2O4 Deposited g-C3N4 Nanosheet for Enhanced Visible Light Photocatalytic Activity. Mater. Sci. Semicond. Process. 2019, 100, 87–97. [Google Scholar] [CrossRef]
  235. Patnaik, S.; Das, K.K.; Mohanty, A.; Parida, K. Enhanced Photo Catalytic Reduction of Cr (VI) over Polymer-Sensitized g-C3N4/ZnFe2O4 and Its Synergism with Phenol Oxidation under Visible Light Irradiation. Catal. Today 2018, 315, 52–66. [Google Scholar] [CrossRef]
  236. Ismael, M.; Wu, Y. A Facile Synthesis Method for Fabrication of LaFeO3/g-C3N4 Nanocomposite as Efficient Visible-Light-Driven Photocatalyst for Photodegradation of RhB and 4-CP. New J. Chem. 2019, 43, 13783–13793. [Google Scholar] [CrossRef]
  237. Khasevani, S.G.; Gholami, M.R. Synthesis of BiOI/ZnFe2O4–Metal–Organic Framework and g-C3N4-Based Nanocomposites for Applications in Photocatalysis. Ind. Eng. Chem. Res. 2019, 58, 9806–9818. [Google Scholar] [CrossRef]
  238. Yu, W.; Chen, J.; Shang, T.; Chen, L.; Gu, L.; Peng, T. Direct Z-Scheme g-C3N4/WO3 Photocatalyst with Atomically Defined Junction for H2 Production. Appl. Catal. B Environ. 2017, 219, 693–704. [Google Scholar] [CrossRef]
  239. Zhang, H.; Feng, Z.; Zhu, Y.; Wu, Y.; Wu, T. Photocatalytic Selective Oxidation of Biomass-Derived 5-Hydroxymethylfurfural to 2, 5-Diformylfuran on WO3/g-C3N4 Composite under Irradiation of Visible Light. J. Photochem. Photobiol. A Chem. 2019, 371, 1–9. [Google Scholar] [CrossRef]
  240. Zhao, Z.; Sun, Y.; Dong, F. Graphitic Carbon Nitride Based Nanocomposites: A Review. Nanoscale 2015, 7, 15–37. [Google Scholar] [CrossRef]
  241. Cheng, C.; Shi, J.; Hu, Y.; Guo, L. WO3/g-C3N4 Composites: One-Pot Preparation and Enhanced Photocatalytic H2 Production under Visible-Light Irradiation. Nanotechnology 2017, 28, 164002. [Google Scholar] [CrossRef] [PubMed]
  242. Yan, H.; Zhu, Z.; Long, Y.; Li, W. Single-Source-Precursor-Assisted Synthesis of Porous WO3/g-C3N4 with Enhanced Photocatalytic Property. Colloids Surfaces A Physicochem. Eng. Asp. 2019, 582, 123857. [Google Scholar] [CrossRef]
  243. Huang, S.; Long, Y.; Ruan, S.; Zeng, Y.-J. Enhanced Photocatalytic CO2 Reduction in Defect-Engineered Z-Scheme WO3–x/g-C3N4 Heterostructures. ACS Omega 2019, 4, 15593–15599. [Google Scholar] [CrossRef] [Green Version]
  244. Zhang, X.; He, S. WOx/g-C3N4 Layered Heterostructures with Controlled Crystallinity towards Superior Photocatalytic Degradation and H2 Generation. Carbon 2020, 156, 488–498. [Google Scholar] [CrossRef]
  245. Han, X.; Xu, D.; An, L.; Hou, C.; Li, Y.; Zhang, Q.; Wang, H. WO3/g-C3N4 Two-Dimensional Composites for Visible-Light Driven Photocatalytic Hydrogen Production. Int. J. Hydrogen Energy 2018, 43, 4845–4855. [Google Scholar] [CrossRef]
  246. Vijayakumar, S.; Vadivel, S. Fiber Optic Ethanol Gas Sensor Based WO3 and WO3/GC3N4 Nanocomposites by a Novel Microwave Technique. Opt. Laser Technol. 2019, 118, 44–51. [Google Scholar] [CrossRef]
  247. Zhao, R.; Li, X.; Su, J.; Gao, X. Preparation of WO3/g-C3N4 Composites and Their Application in Oxidative Desulfurization. Appl. Surf. Sci. 2017, 392, 810–816. [Google Scholar] [CrossRef]
  248. Cadan, F.M.; Ribeiro, C.; Azevedo, E.B. Improving g-C3N4: WO3 Z-Scheme Photocatalytic Performance under Visible Light by Multivariate Optimization of g-C3N4 Synthesis. Appl. Surf. Sci. 2021, 537, 147904. [Google Scholar] [CrossRef]
  249. Singh, J.; Arora, A.; Basu, S. Synthesis of Coral like WO3/g-C3N4 Nanocomposites for the Removal of Hazardous Dyes under Visible Light. J. Alloys Compd. 2019, 808, 151734. [Google Scholar] [CrossRef]
  250. Li, X.; Song, X.; Ma, C.; Cheng, Y.; Shen, D.; Zhang, S.; Liu, W.; Huo, P.; Wang, H. Direct Z-Scheme WO3/Graphitic Carbon Nitride Nanocomposites for the Photoreduction of CO2. ACS Appl. Nano Mater. 2020, 3, 1298–1306. [Google Scholar] [CrossRef]
  251. Chang, F.; Zheng, J.; Wu, F.; Wang, X.; Deng, B. Binary Composites WO3/g-C3N4 in Porous Morphology: Facile Construction, Characterization, and Reinforced Visible Light Photocatalytic Activity. Colloids Surfaces A Physicochem. Eng. Asp. 2019, 563, 11–21. [Google Scholar] [CrossRef]
  252. Fu, J.; Xu, Q.; Low, J.; Jiang, C.; Yu, J. Ultrathin 2D/2D WO3/g-C3N4 Step-Scheme H2-Production Photocatalyst. Appl. Catal. B Environ. 2019, 243, 556–565. [Google Scholar] [CrossRef]
  253. Karimi-Nazarabad, M.; Goharshadi, E.K. Highly Efficient Photocatalytic and Photoelectrocatalytic Activity of Solar Light Driven WO3/g-C3N4 Nanocomposite. Sol. Energy Mater. Sol. Cells 2017, 160, 484–493. [Google Scholar] [CrossRef]
  254. Chai, B.; Liu, C.; Yan, J.; Ren, Z.; Wang, Z. In-Situ Synthesis of WO3 Nanoplates Anchored on g-C3N4 Z-Scheme Photocatalysts for Significantly Enhanced Photocatalytic Activity. Appl. Surf. Sci. 2018, 448, 1–8. [Google Scholar] [CrossRef]
  255. Tahir, B.; Tahir, M.; Nawawi, M.G.M. Highly STable 3D/2D WO3/g-C3N4 Z-Scheme Heterojunction for Stimulating Photocatalytic CO2 Reduction by H2O/H2 to CO and CH4 under Visible Light. J. CO2 Util. 2020, 41, 101270. [Google Scholar] [CrossRef]
  256. Zhang, X.; Wang, X.; Meng, J.; Liu, Y.; Ren, M.; Guo, Y.; Yang, Y. Robust Z-Scheme g-C3N4/WO3 Heterojunction Photocatalysts with Morphology Control of WO3 for Efficient Degradation of Phenolic Pollutants. Sep. Purif. Technol. 2021, 255, 117693. [Google Scholar] [CrossRef]
  257. Borthakur, S.; Basyach, P.; Kalita, L.; Sonowal, K.; Tiwari, A.; Chetia, P.; Saikia, L. Sunlight Assisted Degradation of a Pollutant Dye in Water by a WO3@ g-C3N4 Nanocomposite Catalyst. New J. Chem. 2020, 44, 2947–2960. [Google Scholar] [CrossRef]
  258. Jing, H.; Ou, R.; Yu, H.; Zhao, Y.; Lu, Y.; Huo, M.; Huo, H.; Wang, X. Engineering of g-C3N4 Nanoparticles/WO3 Hollow Microspheres Photocatalyst with Z-Scheme Heterostructure for Boosting Tetracycline Hydrochloride Degradation. Sep. Purif. Technol. 2021, 255, 117646. [Google Scholar] [CrossRef]
  259. Xiao, T.; Tang, Z.; Yang, Y.; Tang, L.; Zhou, Y.; Zou, Z. In Situ Construction of Hierarchical WO3/g-C3N4 Composite Hollow Microspheres as a Z-Scheme Photocatalyst for the Degradation of Antibiotics. Appl. Catal. B Environ. 2018, 220, 417–428. [Google Scholar] [CrossRef]
  260. Liu, X.; Jin, A.; Jia, Y.; Xia, T.; Deng, C.; Zhu, M.; Chen, C.; Chen, X. Synergy of Adsorption and Visible-Light Photocatalytic Degradation of Methylene Blue by a Bifunctional Z-Scheme Heterojunction of WO3/g-C3N4. Appl. Surf. Sci. 2017, 405, 359–371. [Google Scholar] [CrossRef]
  261. Zhuang, H.; Cai, Z.; Xu, W.; Huang, M.; Liu, X. In Situ Construction of WO3/g-C3N4 Composite Photocatalyst with 2D–2D Heterostructure for Enhanced Visible Light Photocatalytic Performance. New J. Chem. 2019, 43, 17416–17422. [Google Scholar] [CrossRef]
  262. Meng, J.; Pei, J.; He, Z.; Wu, S.; Lin, Q.; Wei, X.; Li, J.; Zhang, Z. Facile Synthesis of g-C3N4 Nanosheets Loaded with WO3 Nanoparticles with Enhanced Photocatalytic Performance under Visible Light Irradiation. RSC Adv. 2017, 7, 24097–24104. [Google Scholar] [CrossRef] [Green Version]
  263. Wang, P.; Lu, N.; Su, Y.; Liu, N.; Yu, H.; Li, J.; Wu, Y. Fabrication of WO3@ g-C3N4 with Core@ Shell Nanostructure for Enhanced Photocatalytic Degradation Activity under Visible Light. Appl. Surf. Sci. 2017, 423, 197–204. [Google Scholar] [CrossRef]
  264. Zhao, C.; Ran, F.; Dai, L.; Li, C.; Zheng, C.; Si, C. Cellulose-Assisted Construction of High Surface Area Z-Scheme C-Doped g-C3N4/WO3 for Improved Tetracycline Degradation. Carbohydr. Polym. 2021, 255, 117343. [Google Scholar] [CrossRef]
  265. Mallikarjuna, K.; Kumar, M.K.; Kim, H. Synthesis of Oxygen-Doped-g-C3N4/WO3 Porous Structures for Visible Driven Photocatalytic H2 Production. Phys. E Low-dimensional Syst. Nanostructures 2021, 126, 114428. [Google Scholar]
  266. Sun, L.; Li, B.; Chu, X.; Sun, N.; Qu, Y.; Zhang, X.; Khan, I.; Bai, L.; Jing, L. Synthesis of Si–O-Bridged g-C3N4/WO3 2D-Heterojunctional Nanocomposites as Efficient Photocatalysts for Aerobic Alcohol Oxidation and Mechanism Insight. ACS Sustain. Chem. Eng. 2019, 7, 9916–9927. [Google Scholar] [CrossRef]
  267. Li, H.; Yu, H.; Quan, X.; Chen, S.; Zhang, Y. Uncovering the Key Role of the Fermi Level of the Electron Mediator in a Z-Scheme Photocatalyst by Detecting the Charge Transfer Process of WO3-Metal-GC3N4 (Metal = Cu, Ag, Au). ACS Appl. Mater. Interfaces 2016, 8, 2111–2119. [Google Scholar] [CrossRef]
  268. Chen, J.; Xiao, X.; Wang, Y.; Ye, Z. Ag Nanoparticles Decorated WO3/g-C3N4 2D/2D Heterostructure with Enhanced Photocatalytic Activity for Organic Pollutants Degradation. Appl. Surf. Sci. 2019, 467, 1000–1010. [Google Scholar] [CrossRef]
  269. Fan, G.; Ning, R.; Yan, Z.; Luo, J.; Du, B.; Zhan, J.; Liu, L.; Zhang, J. Double Photoelectron-Transfer Mechanism in Ag− AgCl/WO3/g-C3N4 Photocatalyst with Enhanced Visible-Light Photocatalytic Activity for Trimethoprim Degradation. J. Hazard. Mater. 2021, 403, 123964. [Google Scholar] [CrossRef]
  270. Truc, N.T.T.; Pham, T.-D.; Van Thuan, D.; Tran, D.T.; Nguyen, M.V.; Dang, N.M.; Trang, H.T. Superior Activity of Cu-NiWO4/g-C3N4 Z Direct System for Photocatalytic Decomposition of VOCs in Aerosol under Visible Light. J. Alloys Compd. 2019, 798, 12–18. [Google Scholar] [CrossRef]
  271. Ahmed, K.E.; Kuo, D.-H.; Zeleke, M.A.; Zelekew, O.A.; Abay, A.K. Synthesis of Sn-WO3/g-C3N4 Composites with Surface Activated Oxygen for Visible Light Degradation of Dyes. J. Photochem. Photobiol. A Chem. 2019, 369, 133–141. [Google Scholar] [CrossRef]
  272. Tahir, B.; Tahir, M.; Mohd Nawawi, M.G. Well-Designed 3D/2D/2D WO3/Bt/g-C3N4 Z-Scheme Heterojunction for Tailoring Photocatalytic CO2 Methanation with 2D-Layered Bentonite-Clay as the Electron Moderator under Visible Light. Energy Fuels 2020, 34, 14400–14418. [Google Scholar]
  273. Ouyang, K.; Xu, B.; Yang, C.; Wang, H.; Zhan, P.; Xie, S. Synthesis of a Novel Z-Scheme Ag/WO3/g-C3N4 Nanophotocatalyst for Degradation of Oxytetracycline Hydrochloride under Visible Light. Mater. Sci. Semicond. Process. 2022, 137, 106168. [Google Scholar] [CrossRef]
  274. Qin, Y.; Lu, J.; Meng, F.; Lin, X.; Feng, Y.; Yan, Y.; Meng, M. Rationally Constructing of a Novel 2D/2D WO3/Pt/g-C3N4 Schottky-Ohmic Junction towards Efficient Visible-Light-Driven Photocatalytic Hydrogen Evolution and Mechanism Insight. J. Colloid Interface Sci. 2021, 586, 576–587. [Google Scholar] [CrossRef] [PubMed]
  275. Liu, F.; Feng, S.; Wu, Y.; Lv, X.; Wang, H.; Chen, S.-M.; Hao, Q.; Cao, Y.; Lei, W.; Tong, Z. Label-Free Photoelectrochemical Immunosensor for Aflatoxin B1 Detection Based on the Z-Scheme Heterojunction of g-C3N4/Au/WO3. Biosens. Bioelectron. 2021, 189, 113373. [Google Scholar]
  276. Li, J.; Hao, H.; Zhu, Z. Construction of g-C3N4-WO3-Bi2WO6 Double Z-Scheme System with Enhanced Photoelectrochemical Performance. Mater. Lett. 2016, 168, 180–183. [Google Scholar] [CrossRef]
  277. Sadiq, M.M.J.; Shenoy, U.S.; Bhat, D.K. Synthesis of BaWO4/NRGO–g-C3N4 Nanocomposites with Excellent Multifunctional Catalytic Performance via Microwave Approach. Front. Mater. Sci. 2018, 12, 247–263. [Google Scholar] [CrossRef]
  278. Huang, R.; Gu, X.; Sun, W.; Chen, L.; Du, Q.; Guo, X.; Li, J.; Zhang, M.; Li, C. In Situ Synthesis of Cu+ Self-Doped CuWO4/g-C3N4 Heterogeneous Fenton-like Catalysts: The Key Role of Cu+ in Enhancing Catalytic Performance. Sep. Purif. Technol. 2020, 250, 117174. [Google Scholar] [CrossRef]
  279. Jia, J.; Jiang, C.; Zhang, X.; Li, P.; Xiong, J.; Zhang, Z.; Wu, T.; Wang, Y. Urea-Modified Carbon Quantum Dots as Electron Mediator Decorated g-C3N4/WO3 with Enhanced Visible-Light Photocatalytic Activity and Mechanism Insight. Appl. Surf. Sci. 2019, 495, 143524. [Google Scholar] [CrossRef]
  280. He, K.; Xie, J.; Luo, X.; Wen, J.; Ma, S.; Li, X.; Fang, Y.; Zhang, X. Enhanced Visible Light Photocatalytic H2 Production over Z-Scheme g-C3N4 Nansheets/WO3 Nanorods Nanocomposites Loaded with Ni (OH) x Cocatalysts. Chin. J. Catal. 2017, 38, 240–252. [Google Scholar] [CrossRef]
  281. Jiang, L.; Yuan, X.; Zeng, G.; Liang, J.; Chen, X.; Yu, H.; Wang, H.; Wu, Z.; Zhang, J.; Xiong, T. In-Situ Synthesis of Direct Solid-State Dual Z-Scheme WO3/g-C3N4/Bi2O3 Photocatalyst for the Degradation of Refractory Pollutant. Appl. Catal. B Environ. 2018, 227, 376–385. [Google Scholar] [CrossRef]
  282. Ali, S.; Humayun, M.; Pi, W.; Yuan, Y.; Wang, M.; Khan, A.; Yue, P.; Shu, L.; Zheng, Z.; Fu, Q. Fabrication of BiFeO3-g-C3N4–WO3 Z-Scheme Heterojunction as Highly Efficient Visible-Light Photocatalyst for Water Reduction and 2, 4-Dichlorophenol Degradation: Insight Mechanism. J. Hazard. Mater. 2020, 397, 122708. [Google Scholar] [CrossRef]
  283. Bilal Tahir, M.; Nadeem Riaz, K.; Asiri, A.M. Boosting the Performance of Visible Light-driven WO3/g-C3N4 Anchored with BiVO4 Nanoparticles for Photocatalytic Hydrogen Evolution. Int. J. Energy Res. 2019, 43, 5747–5758. [Google Scholar] [CrossRef]
  284. Beyhaqi, A.; Zeng, Q.; Chang, S.; Wang, M.; Azimi, S.M.T.; Hu, C. Construction of g-C3N4/WO3/MoS2 Ternary Nanocomposite with Enhanced Charge Separation and Collection for Efficient Wastewater Treatment under Visible Light. Chemosphere 2020, 247, 125784. [Google Scholar] [CrossRef]
  285. Cai, Z.; Huang, L.; Quan, X.; Zhao, Z.; Shi, Y.; Puma, G.L. Acetate Production from Inorganic Carbon (HCO3) in Photo-Assisted Biocathode Microbial Electrosynthesis Systems Using WO3/MoO3/g-C3N4 Heterojunctions and Serratia Marcescens Species. Appl. Catal. B Environ. 2020, 267, 118611. [Google Scholar] [CrossRef]
  286. Ibrahim, Y.O.; Gondal, M.A. Visible-Light-Driven Photocatalytic Performance of a Z-Scheme Based TiO2/WO3/g-C3N4 Ternary Heterojunctions. Mol. Catal. 2021, 505, 111494. [Google Scholar] [CrossRef]
  287. Zou, Y.; Xie, Y.; Yu, S.; Chen, L.; Cui, W.; Dong, F.; Zhou, Y. SnO2 Quantum Dots Anchored on g-C3N4 for Enhanced Visible-Light Photocatalytic Removal of NO and Toxic NO2 Inhibition. Appl. Surf. Sci. 2019, 496, 143630. [Google Scholar] [CrossRef]
  288. Wang, Q.; Tian, J.; Wei, L.; Liu, Y.; Yang, C. Z-Scheme Heterostructure of Fe-Doped SnO2 Decorated Layered g-C3N4 with Enhanced Photocatalytic Activity under Simulated Solar Light Irradiation. Opt. Mater. 2020, 101, 109769. [Google Scholar] [CrossRef]
  289. He, Y.; Zhang, L.; Fan, M.; Wang, X.; Walbridge, M.L.; Nong, Q.; Wu, Y.; Zhao, L. Z-Scheme SnO2−x/g-C3N4 Composite as an Efficient Photocatalyst for Dye Degradation and Photocatalytic CO2 Reduction. Sol. Energy Mater. Sol. Cells 2015, 137, 175–184. [Google Scholar] [CrossRef]
  290. Cao, J.; Qin, C.; Wang, Y.; Zhang, H.; Zhang, B.; Gong, Y.; Wang, X.; Sun, G.; Bala, H.; Zhang, Z. Synthesis of g-C3N4 Nanosheet Modified SnO2 Composites with Improved Performance for Ethanol Gas Sensing. RSC Adv. 2017, 7, 25504–25511. [Google Scholar] [CrossRef] [Green Version]
  291. Wang, X.; He, Y.; Xu, L.; Xia, Y.; Gang, R. SnO2 Particles as Efficient Photocatalysts for Organic Dye Degradation Grown In-Situ on g-C3N4 Nanosheets by Microwave-Assisted Hydrothermal Method. Mater. Sci. Semicond. Process. 2021, 121, 105298. [Google Scholar] [CrossRef]
  292. Zhu, K.; Lv, Y.; Liu, J.; Wang, W.; Wang, C.; Li, S.; Wang, P.; Zhang, M.; Meng, A.; Li, Z. Facile Fabrication of g-C3N4/SnO2 Composites and Ball Milling Treatment for Enhanced Photocatalytic Performance. J. Alloys Compd. 2019, 802, 13–18. [Google Scholar] [CrossRef]
  293. Seza, A.; Soleimani, F.; Naseri, N.; Soltaninejad, M.; Montazeri, S.M.; Sadrnezhaad, S.K.; Mohammadi, M.R.; Moghadam, H.A.; Forouzandeh, M.; Amin, M.H. Novel Microwave-Assisted Synthesis of Porous g-C3N4/SnO2 Nanocomposite for Solar Water-Splitting. Appl. Surf. Sci. 2018, 440, 153–161. [Google Scholar] [CrossRef]
  294. Bian, H.; Wang, A.; Li, Z.; Li, Z.; Diao, Y.; Lu, J.; Li, Y.Y. g-C3N4-Modified Water-Crystallized Mesoporous SnO2 for Enhanced Photoelectrochemical Properties. Part. Part. Syst. Charact. 2018, 35, 1800155. [Google Scholar] [CrossRef]
  295. Ji, H.; Fan, Y.; Yan, J.; Xu, Y.; She, X.; Gu, J.; Fei, T.; Xu, H.; Li, H. Construction of SnO2/Graphene-like g-C3N4 with Enhanced Visible Light Photocatalytic Activity. RSC Adv. 2017, 7, 36101–36111. [Google Scholar] [CrossRef] [Green Version]
  296. Zang, Y.; Li, L.; Li, X.; Lin, R.; Li, G. Synergistic Collaboration of g-C3N4/SnO2 Composites for Enhanced Visible-Light Photocatalytic Activity. Chem. Eng. J. 2014, 246, 277–286. [Google Scholar] [CrossRef]
  297. Zhang, T.; Chang, F.; Qi, Y.; Zhang, X.; Yang, J.; Liu, X.; Li, S. A Facile One-Pot and Alkali-Free Synthetic Procedure for Binary SnO2/g-C3N4 Composites with Enhanced Photocatalytic Behavior. Mater. Sci. Semicond. Process. 2020, 115, 105112. [Google Scholar] [CrossRef]
  298. Peng, F.; Ni, Y.; Zhou, Q.; Kou, J.; Lu, C.; Xu, Z. New g-C3N4 Based Photocatalytic Cement with Enhanced Visible-Light Photocatalytic Activity by Constructing Muscovite Sheet/SnO2 Structures. Constr. Build. Mater. 2018, 179, 315–325. [Google Scholar] [CrossRef]
  299. Babu, B.; Cho, M.; Byon, C.; Shim, J. Sunlight-Driven Photocatalytic Activity of SnO2 QDs-g-C3N4 Nanolayers. Mater. Lett. 2018, 212, 327–331. [Google Scholar] [CrossRef]
  300. Shen, H.; Zhao, X.; Duan, L.; Liu, R.; Li, H. Enhanced Visible Light Photocatalytic Activity in SnO2@ g-C3N4 Core-Shell Structures. Mater. Sci. Eng. B 2017, 218, 23–30. [Google Scholar] [CrossRef]
  301. Wang, X.; Ren, P. Flower-like SnO2/g-C3N4 Heterojunctions: The Face-to-Face Contact Interface and Improved Photocatalytic Properties. Adv. Powder Technol. 2018, 29, 1153–1157. [Google Scholar] [CrossRef]
  302. Singh, J.; Kumari, P.; Basu, S. Degradation of Toxic Industrial Dyes Using SnO2/g-C3N4 Nanocomposites: Role of Mass Ratio on Photocatalytic Activity. J. Photochem. Photobiol. A Chem. 2019, 371, 136–143. [Google Scholar] [CrossRef]
  303. Chen, X.; Zhou, B.; Yang, S.; Wu, H.; Wu, Y.; Wu, L.; Pan, J.; Xiong, X. In Situ Construction of an SnO2/g-C3N4 Heterojunction for Enhanced Visible-Light Photocatalytic Activity. Rsc Adv. 2015, 5, 68953–68963. [Google Scholar] [CrossRef]
  304. Yang, L.; Huang, J.; Shi, L.; Cao, L.; Liu, H.; Liu, Y.; Li, Y.; Song, H.; Jie, Y.; Ye, J. Sb Doped SnO2-Decorated Porous g-C3N4 Nanosheet Heterostructures with Enhanced Photocatalytic Activities under Visible Light Irradiation. Appl. Catal. B Environ. 2018, 221, 670–680. [Google Scholar] [CrossRef]
  305. Raziq, F.; Qu, Y.; Humayun, M.; Zada, A.; Yu, H.; Jing, L. Synthesis of SnO2/BP Codoped g-C3N4 Nanocomposites as Efficient Cocatalyst-Free Visible-Light Photocatalysts for CO2 Conversion and Pollutant Degradation. Appl. Catal. B Environ. 2017, 201, 486–494. [Google Scholar] [CrossRef]
  306. Jourshabani, M.; Shariatinia, Z.; Badiei, A. In Situ Fabrication of SnO2/S-Doped g-C3N4 Nanocomposites and Improved Visible Light Driven Photodegradation of Methylene Blue. J. Mol. Liq. 2017, 248, 688–702. [Google Scholar] [CrossRef]
  307. Guo, W.; Huang, L.; Zhang, J.; He, Y.; Zeng, W. Ni-Doped SnO2/g-C3N4 Nanocomposite with Enhanced Gas Sensing Performance for the Eff; Ective Detection of Acetone in Diabetes Diagnosis. Sens. Actuators B Chem. 2021, 334, 129666. [Google Scholar] [CrossRef]
  308. Asaithambi, S.; Sakthivel, P.; Karuppaiah, M.; Yuvakkumar, R.; Velauthapillai, D.; Ahamad, T.; Khan, M.A.M.; Mohammed, M.K.A.; Vijayaprabhu, N.; Ravi, G. The Bifunctional Performance Analysis of Synthesized Ce Doped SnO2/g-C3N4 Composites for Asymmetric Supercapacitor and Visible Light Photocatalytic Applications. J. Alloys Compd. 2021, 866, 158807. [Google Scholar] [CrossRef]
  309. Zada, A.; Humayun, M.; Raziq, F.; Zhang, X.; Qu, Y.; Bai, L.; Qin, C.; Jing, L.; Fu, H. Exceptional Visible-light-driven Cocatalyst-free Photocatalytic Activity of g-C3N4 by Well Designed Nanocomposites with Plasmonic Au and SnO2. Adv. Energy Mater. 2016, 6, 1601190. [Google Scholar] [CrossRef]
  310. Yuan, F.; Sun, Z.; Li, C.; Tan, Y.; Zhang, X.; Zheng, S. Multi-Component Design and in-Situ Synthesis of Visible-Light-Driven SnO2/g-C3N4/Diatomite Composite for High-Efficient Photoreduction of Cr (VI) with the Aid of Citric Acid. J. Hazard. Mater. 2020, 396, 122694. [Google Scholar] [CrossRef]
  311. Babu, B.; Koutavarapu, R.; Shim, J.; Yoo, K. Enhanced Visible-Light-Driven Photoelectrochemical and Photocatalytic Performance of Au-SnO2 Quantum Dot-Anchored g-C3N4 Nanosheets. Sep. Purif. Technol. 2020, 240, 116652. [Google Scholar] [CrossRef]
  312. Peng, L.; Zheng, R.; Feng, D.; Yu, H.; Dong, X. Synthesis of Eco-Friendly Porous g-C3N4/SiO2/SnO2 Composite with Excellent Visible-Light Responsive Photocatalysis. Arab. J. Chem. 2020, 13, 4275–4285. [Google Scholar] [CrossRef]
  313. Wu, J.; Zhang, Y.; Zhou, J.; Wang, K.; Zheng, Y.-Z.; Tao, X. Uniformly Assembling N-Type Metal Oxide Nanostructures (TiO2 Nanoparticles and SnO2 Nanowires) onto P Doped g-C3N4 Nanosheets for Efficient Photocatalytic Water Splitting. Appl. Catal. B Environ. 2020, 278, 119301. [Google Scholar] [CrossRef]
  314. Faraji, M.; Mohaghegh, N.; Abedini, A. Ternary Composite of TiO2 Nanotubes/Ti Plates Modified by g-C3N4 and SnO2 with Enhanced Photocatalytic Activity for Enhancing Antibacterial and Photocatalytic Activity. J. Photochem. Photobiol. B Biol. 2018, 178, 124–132. [Google Scholar] [CrossRef] [PubMed]
  315. Liu, X.-P.; Huang, B.; Mao, C.-J.; Chen, J.-S.; Jin, B.-K. Electrochemiluminescence Aptasensor for Lincomycin Antigen Detection by Using a SnO2/Chitosan/g-C3N4 Nanocomposite. Talanta 2021, 233, 122546. [Google Scholar] [CrossRef]
  316. Ali, G.; Zaidi, S.J.A.; Basit, M.A.; Park, T.J. Synergetic Performance of Systematically Designed g-C3N4/RGO/SnO2 Nanocomposite for Photodegradation of Rhodamine-B Dye. Appl. Surf. Sci. 2021, 570, 151140. [Google Scholar] [CrossRef]
  317. Seo, Y.J.; Das, P.K.; Arunachalam, M.; Ahn, K.-S.; Ha, J.-S.; Kang, S.H. Drawing the Distinguished Graphite Carbon Nitride (g-C3N4) on SnO2 Nanoflake Film for Solar Water Oxidation. Int. J. Hydrogen Energy 2020, 45, 22567–22575. [Google Scholar] [CrossRef]
  318. Wu, H.; Yu, S.; Wang, Y.; Han, J.; Wang, L.; Song, N.; Dong, H.; Li, C. A Facile One-Step Strategy to Construct 0D/2D SnO2/g-C3N4 Heterojunction Photocatalyst for High-Efficiency Hydrogen Production Performance from Water Splitting. Int. J. Hydrogen Energy 2020, 45, 30142–30152. [Google Scholar] [CrossRef]
  319. Versaci, D.; Amici, J.; Francia, C.; Bodoardo, S. Simple Approach Using g-C3N4 to Enable SnO2 Anode High Rate Performance for Li Ion Battery. Solid State Ionics 2020, 346, 115210. [Google Scholar] [CrossRef]
  320. Tran, H.H.; Nguyen, P.H.; Le, M.L.P.; Kim, S.-J.; Vo, V. SnO2 Nanosheets/Graphite Oxide/g-C3N4 Composite as Enhanced Performance Anode Material for Lithium Ion Batteries. Chem. Phys. Lett. 2019, 715, 284–292. [Google Scholar] [CrossRef]
  321. Cao, J.; Qin, C.; Wang, Y.; Zhang, B.; Gong, Y.; Zhang, H.; Sun, G.; Bala, H.; Zhang, Z. Calcination Method Synthesis of SnO2/g-C3N4 Composites for a High-Performance Ethanol Gas Sensing Application. Nanomaterials 2017, 7, 98. [Google Scholar] [CrossRef] [Green Version]
  322. Akhundi, A.; Habibi-Yangjeh, A. A Simple Large-Scale Method for Preparation of g-C3N4/SnO2 Nanocomposite as Visible-Light-Driven Photocatalyst for Degradation of an Organic Pollutant. Mater. Express 2015, 5, 309–318. [Google Scholar] [CrossRef]
  323. Liu, X.; Liu, X. SnO2 Nanocrystallines Decorated g-C3N4 Composites with Enhanced Visible-Light Photocatalytic Activity. Integr. Ferroelectr. 2019, 197, 121–132. [Google Scholar] [CrossRef]
  324. Li, Q.; He, Y.; Peng, R. One-Step Synthesis of SnO 2 Nanoparticles-Loaded Graphitic Carbon Nitride and Their Application in Thermal Decomposition of Ammonium Perchlorate. New J. Chem. 2015, 39, 8703–8707. [Google Scholar] [CrossRef]
  325. Liu, Q.; Fan, C.; Tang, H.; Sun, X.; Yang, J.; Cheng, X. One-Pot Synthesis of g-C3N4/V2O5 Composites for Visible Light-Driven Photocatalytic Activity. Appl. Surf. Sci. 2015, 358, 188–195. [Google Scholar] [CrossRef]
  326. Jayaraman, T.; Raja, S.A.; Priya, A.; Jagannathan, M.; Ashokkumar, M. Synthesis of a Visible-Light Active V2O5–g-C3N4 Heterojunction as an Efficient Photocatalytic and Photoelectrochemical Material. New J. Chem. 2015, 39, 1367–1374. [Google Scholar] [CrossRef]
  327. Jin, C.; Wang, M.; Li, Z.; Kang, J.; Zhao, Y.; Han, J.; Wu, Z. Two Dimensional Co3O4/g-C3N4 Z-Scheme Heterojunction: Mechanism Insight into Enhanced Peroxymonosulfate-Mediated Visible Light Photocatalytic Performance. Chem. Eng. J. 2020, 398, 125569. [Google Scholar] [CrossRef]
  328. Wu, H.; Li, C.; Che, H.; Hu, H.; Hu, W.; Liu, C.; Ai, J.; Dong, H. Decoration of Mesoporous Co3O4 Nanospheres Assembled by Monocrystal Nanodots on g-C3N4 to Construct Z-Scheme System for Improving Photocatalytic Performance. Appl. Surf. Sci. 2018, 440, 308–319. [Google Scholar] [CrossRef]
  329. Qiao, Q.; Yang, K.; Ma, L.-L.; Huang, W.-Q.; Zhou, B.-X.; Pan, A.; Hu, W.; Fan, X.; Huang, G.-F. Facile in Situ Construction of Mediator-Free Direct Z-Scheme g-C3N4/CeO2 Heterojunctions with Highly Efficient Photocatalytic Activity. J. Phys. D Appl. Phys. 2018, 51, 275302. [Google Scholar] [CrossRef]
  330. Huang, L.; Li, Y.; Xu, H.; Xu, Y.; Xia, J.; Wang, K.; Li, H.; Cheng, X. Synthesis and Characterization of CeO2/g-C3N4 Composites with Enhanced Visible-Light Photocatatalytic Activity. Rsc Adv. 2013, 3, 22269–22279. [Google Scholar] [CrossRef]
  331. Zou, W.; Shao, Y.; Pu, Y.; Luo, Y.; Sun, J.; Ma, K.; Tang, C.; Gao, F.; Dong, L. Enhanced Visible Light Photocatalytic Hydrogen Evolution via Cubic CeO2 Hybridized g-C3N4 Composite. Appl. Catal. B Environ. 2017, 218, 51–59. [Google Scholar] [CrossRef]
  332. Li, Y.; Wu, S.; Huang, L.; Xu, H.; Zhang, R.; Qu, M.; Gao, Q.; Li, H. g-C3N4 Modified Bi2O3 Composites with Enhanced Visible-Light Photocatalytic Activity. J. Phys. Chem. Solids 2015, 76, 112–119. [Google Scholar] [CrossRef]
  333. Zhang, J.; Hu, Y.; Jiang, X.; Chen, S.; Meng, S.; Fu, X. Design of a Direct Z-Scheme Photocatalyst: Preparation and Characterization of Bi2O3/g-C3N4 with High Visible Light Activity. J. Hazard. Mater. 2014, 280, 713–722. [Google Scholar] [CrossRef] [PubMed]
  334. Wang, D.; Yu, X.; Feng, Q.; Lin, X.; Huang, Y.; Huang, X.; Li, X.; Chen, K.; Zhao, B.; Zhang, Z. In-Situ Growth of β-Bi2O3 Nanosheets on g-C3N4 to Construct Direct Z-Scheme Heterojunction with Enhanced Photocatalytic Activities. J. Alloys Compd. 2021, 859, 157795. [Google Scholar] [CrossRef]
  335. Wang, X.; Liu, C.; Li, X.; Li, F.; Li, Y.; Zhao, J.; Liu, R. Construction of g-C3N4/Al2O3 Hybrids via in-Situ Acidification and Exfoliation with Enhanced Photocatalytic Activity. Appl. Surf. Sci. 2017, 394, 340–350. [Google Scholar] [CrossRef]
  336. Yang, Z.-L.; Zhang, Z.-Y.; Fan, W.-L.; Hu, C.; Zhang, L.; Qi, J.-J. High-Performance g-C3N4 Added Carbon-Based Perovskite Solar Cells Insulated by Al2O3 Layer. Sol. Energy 2019, 193, 859–865. [Google Scholar]
  337. Shi, J.-W.; Zou, Y.; Cheng, L.; Ma, D.; Sun, D.; Mao, S.; Sun, L.; He, C.; Wang, Z. In-Situ Phosphating to Synthesize Ni2P Decorated NiO/g-C3N4 Pn Junction for Enhanced Photocatalytic Hydrogen Production. Chem. Eng. J. 2019, 378, 122161. [Google Scholar] [CrossRef]
  338. Li, F.; Zhao, Y.; Wang, Q.; Wang, X.; Hao, Y.; Liu, R.; Zhao, D. Enhanced Visible-Light Photocatalytic Activity of Active Al2O3/g-C3N4 Heterojunctions Synthesized via Surface Hydroxyl Modification. J. Hazard. Mater. 2015, 283, 371–381. [Google Scholar] [CrossRef] [PubMed]
  339. Tzvetkov, G.; Tsvetkov, M.; Spassov, T. Ammonia-Evaporation-Induced Construction of Three-Dimensional NiO/g-C3N4 Composite with Enhanced Adsorption and Visible Light-Driven Photocatalytic Performance. Superlattices Microstruct. 2018, 119, 122–133. [Google Scholar] [CrossRef]
  340. Sumathi, M.; Prakasam, A.; Anbarasan, P.M. Fabrication of Hexagonal Disc Shaped Nanoparticles g-C3N4/NiO Heterostructured Nanocomposites for Efficient Visible Light Photocatalytic Performance. J. Clust. Sci. 2019, 30, 757–766. [Google Scholar] [CrossRef]
  341. Huang, L.; Xu, H.; Zhang, R.; Cheng, X.; Xia, J.; Xu, Y.; Li, H. Synthesis and Characterization of g-C3N4/MoO3 Photocatalyst with Improved Visible-Light Photoactivity. Appl. Surf. Sci. 2013, 283, 25–32. [Google Scholar] [CrossRef]
  342. Zhang, X.; Yi, J.; Chen, H.; Mao, M.; Liu, L.; She, X.; Ji, H.; Wu, X.; Yuan, S.; Xu, H. Construction of a Few-Layer g-C3N4/α-MoO3 Nanoneedles All-Solid-State Z-Scheme Photocatalytic System for Photocatalytic Degradation. J. Energy Chem. 2019, 29, 65–71. [Google Scholar] [CrossRef] [Green Version]
  343. Liu, L.; Qi, Y.; Hu, J.; An, W.; Lin, S.; Liang, Y.; Cui, W. Stable Cu2O@ g-C3N4 Core@ Shell Nanostructures: Efficient Visible-Light Photocatalytic Hydrogen Evolution. Mater. Lett. 2015, 158, 278–281. [Google Scholar] [CrossRef]
  344. Han, S.; Hu, X.; Yang, W.; Qian, Q.; Fang, X.; Zhu, Y. Constructing the Band Alignment of Graphitic Carbon Nitride (g-C3N4)/Copper (I) Oxide (Cu2O) Composites by Adjusting the Contact Facet for Superior Photocatalytic Activity. ACS Appl. Energy Mater. 2018, 2, 1803–1811. [Google Scholar] [CrossRef]
  345. Liu, H.; Zhu, X.; Han, R.; Dai, Y.; Sun, Y.; Lin, Y.; Gao, D.; Wang, X.; Luo, C. Study on the Internal Electric Field in the Cu2O/g-C3N4 p–n Heterojunction Structure for Enhancing Visible Light Photocatalytic Activity. New J. Chem. 2020, 44, 1795–1805. [Google Scholar] [CrossRef]
  346. Singh, P.K.; Bhardiya, S.R.; Asati, A.; Rai, V.K.; Singh, M.; Rai, A. Cu/Cu2O@ g-C3N4: Recyclable Photocatalyst under Visible Light to Access 2-Aryl-/Benzimidazoles/Benzothiazoles in Water. ChemistrySelect 2020, 5, 14270–14275. [Google Scholar] [CrossRef]
  347. Liang, S.; Zhou, Y.; Cai, Z.; She, C. Preparation of Porous g-C3N4/Ag/Cu2O: A New Composite with Enhanced Visible-light Photocatalytic Activity. Appl. Organomet. Chem. 2016, 30, 932–938. [Google Scholar] [CrossRef]
  348. Chaudhary, P.; Ingole, P.P. In-Situ Solid-State Synthesis of 2D/2D Interface between Ni/NiO Hexagonal Nanosheets Supported on g-C3N4 for Enhanced Photo-Electrochemical Water Splitting. Int. J. Hydrogen Energy 2020, 45, 16060–16070. [Google Scholar] [CrossRef]
  349. Zhou, C.; Liu, Z.; Fang, L.; Guo, Y.; Feng, Y.; Yang, M. Kinetic and Mechanistic Study of Rhodamine B Degradation by H2O2 and Cu/Al2O3/g-C3N4 Composite. Catalysts 2020, 10, 317. [Google Scholar] [CrossRef] [Green Version]
  350. Bao, Y.; Chen, K. A Novel Z-Scheme Visible Light Driven Cu2O/Cu/g-C3N4 Photocatalyst Using Metallic Copper as a Charge Transfer Mediator. Mol. Catal. 2017, 432, 187–195. [Google Scholar] [CrossRef]
  351. Zhang, P.; Wang, T.; Zeng, H. Design of Cu-Cu2O/g-C3N4 Nanocomponent Photocatalysts for Hydrogen Evolution under Visible Light Irradiation Using Water-Soluble Erythrosin B Dye Sensitization. Appl. Surf. Sci. 2017, 391, 404–414. [Google Scholar] [CrossRef]
  352. Liu, C.; Mao, D.; Pan, J.; Qian, J.; Zhang, W.; Chen, F.; Chen, Z.; Song, Y. Fabrication of Highly Efficient Heterostructured Ag-CeO2/g-C3N4 Hybrid Photocatalyst with Enhanced Visible-Light Photocatalytic Activity. J. Rare Earths 2019, 37, 1269–1278. [Google Scholar] [CrossRef]
  353. Pinheiro, D.; Devi, K.R.S.; Jose, A.; Karthik, K.; Sugunan, S.; Mohan, M.K. Experimental Design for Optimization of 4-Nitrophenol Reduction by Green Synthesized CeO2/g-C3N4/Ag Catalyst Using Response Surface Methodology. J. Rare Earths 2020, 38, 1171–1177. [Google Scholar] [CrossRef]
  354. Yin, H.Y.; Zheng, Y.F.; Wang, L. Au/CeO2/g-C3N4 Nanocomposite Modified Electrode as Electrochemical Sensor for the Determination of Phenol. J. Nanosci. Nanotechnol. 2020, 20, 5539–5545. [Google Scholar] [CrossRef]
  355. Tian, N.; Huang, H.; Guo, Y.; He, Y.; Zhang, Y. A g-C3N4/Bi2O2CO3 Composite with High Visible-Light-Driven Photocatalytic Activity for Rhodamine B Degradation. Appl. Surf. Sci. 2014, 322, 249–254. [Google Scholar] [CrossRef]
  356. Mohammad, A.; Karim, M.R.; Khan, M.E.; Khan, M.M.; Cho, M.H. Biofilm-Assisted Fabrication of Ag@ SnO2-g-C3N4 Nanostructures for Visible Light-Induced Photocatalysis and Photoelectrochemical Performance. J. Phys. Chem. C 2019, 123, 20936–20948. [Google Scholar] [CrossRef]
  357. Saravanakumar, K.; Karthik, R.; Chen, S.-M.; Kumar, J.V.; Prakash, K.; Muthuraj, V. Construction of Novel Pd/CeO2/g-C3N4 Nanocomposites as Efficient Visible-Light Photocatalysts for Hexavalent Chromium Detoxification. J. Colloid Interface Sci. 2017, 504, 514–526. [Google Scholar] [CrossRef] [PubMed]
  358. Xie, Q.; He, W.; Liu, S.; Li, C.; Zhang, J.; Wong, P.K. Bifunctional S-Scheme g-C3N4/Bi/BiVO4 Hybrid Photocatalysts toward Artificial Carbon Cycling. Chin. J. Catal. 2020, 41, 140–153. [Google Scholar] [CrossRef]
  359. Han, C.; Zhang, R.; Ye, Y.; Wang, L.; Ma, Z.; Su, F.; Xie, H.; Zhou, Y.; Wong, P.K.; Ye, L. Chainmail Co-Catalyst of NiO Shell-Encapsulated Ni for Improving Photocatalytic CO2 Reduction over GC3N4. J. Mater. Chem. A 2019, 7, 9726–9735. [Google Scholar] [CrossRef]
  360. Zhao, X.; Guan, J.; Li, J.; Li, X.; Wang, H.; Huo, P.; Yan, Y. CeO2/3D g-C3N4 Heterojunction Deposited with Pt Cocatalyst for Enhanced Photocatalytic CO2 Reduction. Appl. Surf. Sci. 2021, 537, 147891. [Google Scholar] [CrossRef]
  361. Karimi, M.A.; Atashkadi, M.; Ranjbar, M.; Habibi-Yangjeh, A. Novel Visible-Light-Driven Photocatalyst of NiO/Cd/g-C3N4 for Enhanced Degradation of Methylene Blue. Arab. J. Chem. 2020, 13, 5810–5820. [Google Scholar] [CrossRef]
  362. Jia, Z.; Lyu, F.; Zhang, L.C.; Zeng, S.; Liang, S.X.; Li, Y.Y.; Lu, J. Pt Nanoparticles Decorated Heterostructured g-C3N4/Bi2MoO6 Microplates with Highly Enhanced Photocatalytic Activities under Visible Light. Sci. Rep. 2019, 9, 1–13. [Google Scholar] [CrossRef] [Green Version]
  363. Guo, Q.; Zhang, C.; Zhang, C.; Xin, S.; Zhang, P.; Shi, Q.; Zhang, D.; You, Y. Co3O4 Modified Ag/g-C3N4 Composite as a Bifunctional Cathode for Lithium-Oxygen Battery. J. Energy Chem. 2020, 41, 185–193. [Google Scholar] [CrossRef] [Green Version]
  364. Xie, Z.; Feng, Y.; Wang, F.; Chen, D.; Zhang, Q.; Zeng, Y.; Lv, W.; Liu, G. Construction of Carbon Dots Modified MoO3/g-C3N4 Z-Scheme Photocatalyst with Enhanced Visible-Light Photocatalytic Activity for the Degradation of Tetracycline. Appl. Catal. B Environ. 2018, 229, 96–104. [Google Scholar] [CrossRef]
  365. Shi, W.; Wang, J.; Yang, S.; Lin, X.; Guo, F.; Shi, J. Fabrication of a Ternary Carbon Dots/CoO/g-C3N4 Nanocomposite Photocatalyst with Enhanced Visible-light-driven Photocatalytic Hydrogen Production. J. Chem. Technol. Biotechnol. 2020, 95, 2129–2138. [Google Scholar] [CrossRef]
  366. Kumar, A.; Sharma, S.K.; Sharma, G.; Naushad, M.; Stadler, F.J. CeO2/g-C3N4/V2O5 Ternary Nano Hetero-Structures Decorated with CQDs for Enhanced Photo-Reduction Capabilities under Different Light Sources: Dual Z-Scheme Mechanism. J. Alloys Compd. 2020, 838, 155692. [Google Scholar] [CrossRef]
  367. Wang, H.; Yang, Y.; Qu, T.; Kang, Z.; Wang, D. Co3O4 and CDots Nanocrystals on g-C3N4 as a Synergetic Catalyst for Oxygen Reduction Reaction. Green Process. Synth. 2015, 4, 411–419. [Google Scholar] [CrossRef]
  368. Shekardasht, M.B.; Givianrad, M.H.; Gharbani, P.; Mirjafary, Z.; Mehrizad, A. Preparation of a Novel Z-Scheme g-C3N4/RGO/Bi2Fe4O9 Nanophotocatalyst for Degradation of Congo Red Dye under Visible Light. Diam. Relat. Mater. 2020, 109, 108008. [Google Scholar] [CrossRef]
  369. Gong, L.; Li, X.; Zhang, Q.; Huang, B.; Yang, Q.; Yang, G.; Liu, Y. Ultrafast and Large-Scale Synthesis of Co3O4 Quantum Dots-C3N4/RGO as an Excellent ORR Electrocatalyst via a Controllable Deflagration Strategy. Appl. Surf. Sci. 2020, 525, 146624. [Google Scholar] [CrossRef]
  370. Feng, Y.; Yan, T.; Wu, T.; Zhang, N.; Yang, Q.; Sun, M.; Yan, L.; Du, B.; Wei, Q. A Label-Free Photoelectrochemical Aptasensing Platform Base on Plasmon Au Coupling with MOF-Derived In2O3@ g-C3N4 Nanoarchitectures for Tetracycline Detection. Sensors Actuators B Chem. 2019, 298, 126817. [Google Scholar] [CrossRef]
  371. Cui, Y.; Nengzi, L.; Gou, J.; Huang, Y.; Li, B.; Cheng, X. Fabrication of Dual Z-Scheme MIL-53 (Fe)/α-Bi2O3/g-C3N4 Ternary Composite with Enhanced Visible Light Photocatalytic Performance. Sep. Purif. Technol. 2020, 232, 115959. [Google Scholar] [CrossRef]
  372. Jourshabani, M.; Shariatinia, Z.; Badiei, A. Synthesis and Characterization of Novel Sm2O3/S-Doped g-C3N4 Nanocomposites with Enhanced Photocatalytic Activities under Visible Light Irradiation. Appl. Surf. Sci. 2018, 427, 375–387. [Google Scholar] [CrossRef]
  373. Lan, Y.; Li, Z.; Xie, W.; Li, D.; Yan, G.; Guo, S.; Pan, C.; Wu, J. In Situ Fabrication of I-Doped Bi2O2CO3/g-C3N4 Heterojunctions for Enhanced Photodegradation Activity under Visible Light. J. Hazard. Mater. 2020, 385, 121622. [Google Scholar] [CrossRef] [PubMed]
  374. Xue, S.; Hou, X.; Xie, W.; Wei, X.; He, D. Dramatic Improvement of Photocatalytic Activity for N-Doped Bi2O3/g-C3N4 Composites. Mater. Lett. 2015, 161, 640–643. [Google Scholar] [CrossRef]
  375. Zhou, P.; Meng, X.; Sun, T. Facile Fabrication of In2O3/S-Doped g-C3N4 Heterojunction Hybrids for Enhanced Visible-Light Photocatalytic Hydrogen Evolution. Mater. Lett. 2020, 261, 127159. [Google Scholar] [CrossRef]
  376. Jin, X.; Guan, Q.; Tian, T.; Li, H.; Han, Y.; Hao, F.; Cui, Y.; Li, W.; Zhu, Y.; Zhang, Y. In2O3/Boron Doped g-C3N4 Heterojunction Catalysts with Remarkably Enhanced Visible-Light Photocatalytic Efficiencies. Appl. Surf. Sci. 2020, 504, 144241. [Google Scholar] [CrossRef]
  377. Liu, J.-L.; Jiang, J.; Zhang, J.-Q.; Chai, Y.-Q.; Xiao, Q.; Yuan, R. The Combination of Ternary Electrochemiluminescence System of g-C3N4 Nanosheet/TEA/Cu@ Cu2O and G-Quadruplex-Driven Regeneration Strategy for Ultrasensitive Bioanalysis. Biosens. Bioelectron. 2020, 152, 112006. [Google Scholar] [CrossRef]
  378. Li, J.; Li, Y.; Zhang, W.; Naraginti, S.; Sivakumar, A.; Zhang, C. Fabrication of Novel Tetrahedral Ag3PO4/g-C3N4/BiVO4 Ternary Composite for Efficient Detoxification of Sulfamethoxazole. Process Saf. Environ. Prot. 2020, 143, 340–347. [Google Scholar] [CrossRef]
  379. Hu, B.; Liu, Y.; Wang, Z.-W.; Song, Y.; Wang, M.; Zhang, Z.; Liu, C.-S. Bimetallic-Organic Framework Derived Porous Co3O4/Fe3O4/C-Loaded g-C3N4 Nanocomposites as Non-Enzymic Electrocatalysis Oxidization toward Ascorbic Acid, Dopamine Acid, and Uric Acid. Appl. Surf. Sci. 2018, 441, 694–707. [Google Scholar] [CrossRef]
  380. Wang, Y.; Yu, D.; Wang, W.; Gao, P.; Zhong, S.; Zhang, L.; Zhao, Q.; Liu, B. Synthesizing Co3O4-BiVO4/g-C3N4 Heterojunction Composites for Superior Photocatalytic Redox Activity. Sep. Purif. Technol. 2020, 239, 116562. [Google Scholar] [CrossRef]
  381. Sudrajat, H.; Hartuti, S. Structural Properties and Catalytic Activity of a Novel Ternary CuO/GC3N4/Bi2O3 Photocatalyst. J. Colloid Interface Sci. 2018, 524, 227–235. [Google Scholar] [CrossRef]
  382. Zhao, H.; Zhang, J.; Zheng, H. Facile Preparation of Dual Z-Scheme Bi2O3/g-C3N4/Ag6Si2O7 Photocatalyst for Highly Efficient Visible-Light Photocatalytic Degradation of Organic Pollutants. Appl. Surf. Sci. 2020, 527, 146901. [Google Scholar] [CrossRef]
  383. Vattikuti, S.V.P.; Police, A.K.R.; Shim, J.; Byon, C. In Situ Fabrication of the Bi2O3–V2O5 Hybrid Embedded with Graphitic Carbon Nitride Nanosheets: Oxygen Vacancies Mediated Enhanced Visible-Light–Driven Photocatalytic Degradation of Organic Pollutants and Hydrogen Evolution. Appl. Surf. Sci. 2018, 447, 740–756. [Google Scholar] [CrossRef]
  384. Wu, S.; Guo, J.; Wang, Y. Bi2O2CO3–Bi2O2(OH)NO3/g-C3N4 Heterojunction as a Visible-Light-Driven Photocatalyst with Enhanced Photogenerated Charge Separation. J. Alloys Compd. 2020, 818, 152852. [Google Scholar] [CrossRef]
  385. Zhong, S.; Zhou, H.; Shen, M.; Yao, Y.; Gao, Q. Rationally Designed a g-C3N4/BiOI/Bi2O2CO3 Composite with Promoted Photocatalytic Activity. J. Alloys Compd. 2021, 853, 157307. [Google Scholar] [CrossRef]
  386. Kumar, A.; Kumar, A.; Sharma, G.; Ala’a, H.; Naushad, M.; Ghfar, A.A.; Guo, C.; Stadler, F.J. Biochar-Templated g-C3N4/Bi2O2CO3/CoFe2O4 Nano-Assembly for Visible and Solar Assisted Photo-Degradation of Paraquat, Nitrophenol Reduction and CO2 Conversion. Chem. Eng. J. 2018, 339, 393–410. [Google Scholar] [CrossRef]
  387. Min, Z.; Wang, X.; Li, Y.; Jiang, J.; Li, J.; Qian, D.; Li, J. A Highly Efficient Visible-Light-Responding Cu2O∓TiO2/g-C3N4 Photocatalyst for Instantaneous Discolorations of Organic Dyes. Mater. Lett. 2017, 193, 18–21. [Google Scholar] [CrossRef]
  388. Muñoz-Batista, M.J.; Nasalevich, M.A.; Savenije, T.J.; Kapteijn, F.; Gascon, J.; Kubacka, A.; Fernández-García, M. Enhancing Promoting Effects in g-C3N4-Mn+/CeO2-TiO2 Ternary Composites: Photo-Handling of Charge Carriers. Appl. Catal. B Environ. 2015, 176, 687–698. [Google Scholar] [CrossRef]
  389. da Silva Veras, T.; Mozer, T.S.; da Silva César, A. Hydrogen: Trends, Production and Characterization of the Main Process Worldwide. Int. J. Hydrogen Energy 2017, 42, 2018–2033. [Google Scholar] [CrossRef]
  390. He, R.; Liang, H.; Li, C.; Bai, J. Enhanced Photocatalytic Hydrogen Production over Co3O4@ g-C3N4 Pn Junction Adhering on One-Dimensional Carbon Fiber. Colloids Surfaces A Physicochem. Eng. Asp. 2020, 586, 124200. [Google Scholar] [CrossRef]
  391. Anandan, S.; Wu, J.J.; Bahnemann, D.; Emeline, A.; Ashokkumar, M. Crumpled Cu2O-g-C3N4 Nanosheets for Hydrogen Evolution Catalysis. Colloids Surfaces A Physicochem. Eng. Asp. 2017, 527, 34–41. [Google Scholar] [CrossRef]
  392. Shi, J.; Zheng, B.; Mao, L.; Cheng, C.; Hu, Y.; Wang, H.; Li, G.; Jing, D.; Liang, X. MoO3/g-C3N4 Z-Scheme (S-Scheme) System Derived from MoS2/Melamine Dual Precursors for Enhanced Photocatalytic H2 Evolution Driven by Visible Light. Int. J. Hydrogen Energy 2021, 46, 2927–2935. [Google Scholar] [CrossRef]
  393. Chebanenko, M.I.; Lobinsky, A.A.; Nevedomskiy, V.N.; Popkov, V.I. NiO-Decorated Graphitic Carbon Nitride toward Electrocatalytic Hydrogen Production from Ethanol. Dalt. Trans. 2020, 49, 12088–12097. [Google Scholar] [CrossRef] [PubMed]
  394. Ma, X.; Zhang, J.; Wang, B.; Li, Q.; Chu, S. Hierarchical Cu2O Foam/g-C3N4 Photocathode for Photoelectrochemical Hydrogen Production. Appl. Surf. Sci. 2018, 427, 907–916. [Google Scholar] [CrossRef]
  395. Paul, A.M.; Sajeev, A.; Nivetha, R.; Gothandapani, K.; Bhardwaj, P.; Govardhan, K.; Raghavan, V.; Jacob, G.; Sellapan, R.; Jeong, S.K. Cuprous Oxide (Cu2O)/Graphitic Carbon Nitride (g-C3N4) Nanocomposites for Electrocatalytic Hydrogen Evolution Reaction. Diam. Relat. Mater. 2020, 107, 107899. [Google Scholar] [CrossRef]
  396. Zhang, M.; Qin, J.; Rajendran, S.; Zhang, X.; Liu, R. Heterostructured D-Ti3C2/TiO2/g-C3N4 Nanocomposites with Enhanced Visible-light Photocatalytic Hydrogen Production Activity. ChemSusChem 2018, 11, 4226–4236. [Google Scholar] [CrossRef] [PubMed]
  397. Wen, P.; Sun, Y.; Li, H.; Liang, Z.; Wu, H.; Zhang, J.; Zeng, H.; Geyer, S.M.; Jiang, L. A Highly Active Three-Dimensional Z-Scheme ZnO/Au/g-C3N4 Photocathode for Efficient Photoelectrochemical Water Splitting. Appl. Catal. B Environ. 2020, 263, 118180. [Google Scholar] [CrossRef]
  398. Wei, X.; Wang, X.; Pu, Y.; Liu, A.; Chen, C.; Zou, W.; Zheng, Y.; Huang, J.; Zhang, Y.; Yang, Y. Facile Ball-Milling Synthesis of CeO2/g-C3N4 Z-Scheme Heterojunction for Synergistic Adsorption and Photodegradation of Methylene Blue: Characteristics, Kinetics, Models, and Mechanisms. Chem. Eng. J. 2020, 470, 127719. [Google Scholar] [CrossRef]
  399. Zhang, S.; Yan, J.; Yang, S.; Xu, Y.; Cai, X.; Li, X.; Zhang, X.; Peng, F.; Fang, Y. Electrodeposition of Cu2O/g-C3N4 Heterojunction Film on an FTO Substrate for Enhancing Visible Light Photoelectrochemical Water Splitting. Chin. J. Catal. 2017, 38, 365–371. [Google Scholar] [CrossRef]
  400. Ji, C.; Yin, S.-N.; Sun, S.; Yang, S. An in Situ Mediator-Free Route to Fabricate Cu2O/g-C3N4 Type-II Heterojunctions for Enhanced Visible-Light Photocatalytic H2 Generation. Appl. Surf. Sci. 2018, 434, 1224–1231. [Google Scholar] [CrossRef]
  401. Liu, L.; Qi, Y.; Hu, J.; Liang, Y.; Cui, W. Efficient Visible-Light Photocatalytic Hydrogen Evolution and Enhanced Photostability of Core@ Shell Cu2O@ g-C3N4 Octahedra. Appl. Surf. Sci. 2015, 351, 1146–1154. [Google Scholar] [CrossRef]
  402. Yan, H.; Yang, H. TiO2–g-C3N4 Composite Materials for Photocatalytic H2 Evolution under Visible Light Irradiation. J. Alloys Compd. 2011, 509, L26–L29. [Google Scholar] [CrossRef]
  403. Marchal, C.; Cottineau, T.; Méndez-Medrano, M.G.; Colbeau-Justin, C.; Caps, V.; Keller, V. Au/TiO2/GC3N4 Nanocomposites for Enhanced Photocatalytic H2 Production from Water under Visible Light Irradiation with Very Low Quantities of Sacrificial Agents. Adv. Energy Mater. 2018, 8, 1702142. [Google Scholar] [CrossRef]
  404. Geng, R.; Yin, J.; Zhou, J.; Jiao, T.; Feng, Y.; Zhang, L.; Chen, Y.; Bai, Z.; Peng, Q. In Situ Construction of Ag/TiO2/g-C3N4 Heterojunction Nanocomposite Based on Hierarchical Co-Assembly with Sustainable Hydrogen Evolution. Nanomaterials 2020, 10, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  405. Pan, J.; You, M.; Chi, C.; Dong, Z.; Wang, B.; Zhu, M.; Zhao, W.; Song, C.; Zheng, Y.; Li, C. The Two Dimension Carbon Quantum Dots Modified Porous g-C3N4/TiO2 Nano-Heterojunctions for Visible Light Hydrogen Production Enhancement. Int. J. Hydrogen Energy 2018, 43, 6586–6593. [Google Scholar] [CrossRef]
  406. Mahala, C.; Sharma, M.D.; Basu, M. ZnO Nanosheets Decorated with Graphite-like Carbon Nitride Quantum Dots as Photoanodes in Photoelectrochemical Water Splitting. ACS Appl. Nano Mater. 2020, 3, 1999–2007. [Google Scholar] [CrossRef]
  407. Qin, H.; Zuo, Y.; Jin, J.; Wang, W.; Xu, Y.; Cui, L.; Dang, H. ZnO Nanorod Arrays Grown on g-C3N4 Micro-Sheets for Enhanced Visible Light Photocatalytic H2 Evolution. RSC Adv. 2019, 9, 24483–24488. [Google Scholar] [CrossRef] [Green Version]
  408. Zhang, Y.; He, S.; Guo, W.; Hu, Y.; Huang, J.; Mulcahy, J.R.; Wei, W.D. Surface-Plasmon-Driven Hot Electron Photochemistry. Chem. Rev. 2017, 118, 2927–2954. [Google Scholar] [CrossRef] [PubMed]
  409. Saha, D.; Gismondi, P.; Kolasinski, K.W.; Shumlas, S.L.; Rangan, S.; Eslami, B.; McConnell, A.; Bui, T.; Cunfer, K. Fabrication of Electrospun Nanofiber Composite of g-C3N4 and Au Nanoparticles as Plasmonic Photocatalyst. Surfaces Interfaces 2021, 26, 101367. [Google Scholar] [CrossRef]
  410. Zhang, X.; Chen, Y.L.; Liu, R.-S.; Tsai, D.P. Plasmonic Photocatalysis. Reports Prog. Phys. 2013, 76, 46401. [Google Scholar] [CrossRef] [Green Version]
  411. Zhao, W.; Dong, Q.; Sun, C.; Xia, D.; Huang, H.; Yang, G.; Wang, G.; Leung, D.Y.C. A Novel Au/g-C3N4 Nanosheets/CeO2 Hollow Nanospheres Plasmonic Heterojunction Photocatalysts for the Photocatalytic Reduction of Hexavalent Chromium and Oxidation of Oxytetracycline Hydrochloride. Chem. Eng. J. 2021, 409, 128185. [Google Scholar] [CrossRef]
  412. Bashiri, R.; Mohamed, N.M.; Suhaimi, N.A.; Shahid, M.U.; Kait, C.F.; Sufian, S.; Khatani, M.; Mumtaz, A. Photoelectrochemical Water Splitting with Tailored TiO2/SrTiO3@ g-C3N4 Heterostructure Nanorod in Photoelectrochemical Cell. Diam. Relat. Mater. 2018, 85, 5–12. [Google Scholar] [CrossRef]
  413. Wang, S.; Zheng, F.; Weng, Y.; Yuan, G.; Fang, L.; You, L. Enhanced Photoelectrochemical Performance by Interface Engineering in Ternary g-C3N4/TiO2/PbTiO3 Films. Adv. Mater. Interfaces 2020, 7, 2000185. [Google Scholar] [CrossRef]
  414. Qu, A.; Xu, X.; Xie, H.; Zhang, Y.; Li, Y.; Wang, J. Effects of Calcining Temperature on Photocatalysis of g-C3N4/TiO2 Composites for Hydrogen Evolution from Water. Mater. Res. Bull. 2016, 80, 167–176. [Google Scholar] [CrossRef]
  415. Jiang, Y.; Sun, Z.; Chen, Q.; Cao, C.; Zhao, Y.; Yang, W.; Zeng, L.; Huang, L. Fabrication of 0D/2D TiO2 Nanodots/g-C3N4 S-Scheme Heterojunction Photocatalyst for Efficient Photocatalytic Overall Water Splitting. Appl. Surf. Sci. 2021, 571, 151287. [Google Scholar] [CrossRef]
  416. Liu, J.; Yan, X.-T.; Qin, X.-S.; Wu, S.-J.; Zhao, H.; Yu, W.-B.; Chen, L.-H.; Li, Y.; Su, B.-L. Light-Assisted Preparation of Heterostructured g-C3N4/ZnO Nanorods Arrays for Enhanced Photocatalytic Hydrogen Performance. Catal. Today 2019, 355, 932–936. [Google Scholar] [CrossRef]
  417. Liu, Y.; Liu, H.; Zhou, H.; Li, T.; Zhang, L. A Z-Scheme Mechanism of N-ZnO/g-C3N4 for Enhanced H2 Evolution and Photocatalytic Degradation. Appl. Surf. Sci. 2019, 466, 133–140. [Google Scholar] [CrossRef]
  418. Gu, Y.; Bao, A.; Zhang, X.; Yan, J.; Du, Q.; Zhang, M.; Qi, X. Facile Fabrication of Sulfur-Doped Cu2O and g-C3N4 with Z-Scheme Structure for Enhanced Photocatalytic Water Splitting Performance. Mater. Chem. Phys. 2021, 266, 124542. [Google Scholar] [CrossRef]
  419. Alhaddad, M.; Navarro, R.M.; Hussein, M.A.; Mohamed, R.M. Bi2O3/g-C3N4 Nanocomposites as Proficient Photocatalysts for Hydrogen Generation from Aqueous Glycerol Solutions beneath Visible Light. Ceram. Int. 2020, 46, 24873–24881. [Google Scholar] [CrossRef]
  420. Ma, R.; Dong, L.; Li, B.; Su, T.; Luo, X.; Qin, Z.; Ji, H. g-C3N4/BiYO3 Composite for Photocatalytic Hydrogen Evolution. ChemistrySelect 2018, 3, 5891–5899. [Google Scholar] [CrossRef]
  421. Xu, J.; Yu, H.; Guo, H. Synthesis and Behaviors of g-C3N4 Coupled with LaxCo3-XO4 Nanocomposite for Improved Photocatalytic Activeity and Stability under Visible Light. Mater. Res. Bull. 2018, 105, 342–348. [Google Scholar] [CrossRef]
  422. Li, Y.; Zhu, S.; Kong, X.; Liang, Y.; Li, Z.; Wu, S.; Chang, C.; Luo, S.; Cui, Z. In Situ Synthesis of a Novel Mn3O4/g-C3N4 Pn Heterostructure Photocatalyst for Water Splitting. J. Colloid Interface Sci. 2021, 586, 778–784. [Google Scholar] [CrossRef] [PubMed]
  423. Song, T.; Zhang, X.; Yang, P. Bifunctional Nitrogen-Doped Carbon Dots in g-C3N4/WOx Heterojunction for Enhanced Photocatalytic Water-Splitting Performance. Langmuir 2021, 37, 4236–4247. [Google Scholar] [CrossRef]
  424. Fu, Y.; Zhu, C.; Wang, H.; Dou, Y.; Shi, W.; Shao, M.; Huang, H.; Liu, Y.; Kang, Z. High-Performance NiO/g-C3N4 Composites for Visible-Light-Driven Photocatalytic Overall Water Splitting. Inorg. Chem. Front. 2018, 5, 1646–1652. [Google Scholar] [CrossRef]
  425. Cao, S.-W.; Liu, X.-F.; Yuan, Y.-P.; Zhang, Z.-Y.; Liao, Y.-S.; Fang, J.; Loo, S.C.J.; Sum, T.C.; Xue, C. Solar-to-Fuels Conversion over In2O3/g-C3N4 Hybrid Photocatalysts. Appl. Catal. B Environ. 2014, 147, 940–946. [Google Scholar] [CrossRef]
  426. Guo, Y.; Chang, B.; Wen, T.; Zhang, S.; Zeng, M.; Hu, N.; Su, Y.; Yang, Z.; Yang, B. A Z-Scheme Photocatalyst for Enhanced Photocatalytic H2 Evolution, Constructed by Growth of 2D Plasmonic MoO3−x Nanoplates onto 2D g-C3N4 Nanosheets. J. Colloid Interface Sci. 2020, 567, 213–223. [Google Scholar] [CrossRef]
  427. Hou, H.L.; Gao, F.; Wang, L.; Shang, M.; Yang, Z.; Zheng, J.; Yang, W. Superior Ternary Hybrid Photocatalysts of TiO2/WO3/GC3N4 Thoroughly Mesoporous Nanofibers for Visible-Lightdriven Hydrogen Evolution. J. Mater. Chem. A 2016, 4, 6276–6281. [Google Scholar] [CrossRef]
  428. Hieu, V.Q.; Lam, T.C.; Khan, A.; Vo, T.-T.T.; Nguyen, T.-Q.; Doan, V.D.; Tran, V.A. TiO2/Ti3C2/g-C3N4 Ternary Heterojunction for Photocatalytic Hydrogen Evolution. Chemosphere 2021, 285, 131429. [Google Scholar] [CrossRef]
  429. Nie, N.; Zhang, L.; Fu, J.; Cheng, B.; Yu, J. Self-Assembled Hierarchical Direct Z-Scheme g-C3N4/ZnO Microspheres with Enhanced Photocatalytic CO2 Reduction Performance. Appl. Surf. Sci. 2018, 441, 12–22. [Google Scholar] [CrossRef]
  430. Mulik, B.B.; Bankar, B.D.; Munde, A.V.; Chavan, P.P.; Biradar, A.V.; Sathe, B.R. Electrocatalytic and Catalytic CO2 Hydrogenation on ZnO/g-C3N4 Hybrid Nanoelectrodes. Appl. Surf. Sci. 2021, 538, 148120. [Google Scholar] [CrossRef]
  431. Wang, H.; Li, H.; Chen, Z.; Li, J.; Li, X.; Huo, P.; Wang, Q. TiO2 Modified g-C3N4 with Enhanced Photocatalytic CO2 Reduction Performance. Solid State Sci. 2020, 100, 106099. [Google Scholar] [CrossRef]
  432. Truc, N.T.T.; Bach, L.G.; Hanh, N.T.; Pham, T.-D.; Le Chi, N.T.P.; Tran, D.T.; Nguyen, M.V. The Superior Photocatalytic Activity of Nb Doped TiO2/g-C3N4 Direct Z-Scheme System for Efficient Conversion of CO2 into Valuable Fuels. J. Colloid Interface Sci. 2019, 540, 1–8. [Google Scholar] [CrossRef] [PubMed]
  433. Guo, Q.; Fu, L.; Yan, T.; Tian, W.; Ma, D.; Li, J.; Jiang, Y.; Wang, X. Improved Photocatalytic Activity of Porous ZnO Nanosheets by Thermal Deposition Graphene-like g-C3N4 for CO2 Reduction with H2O Vapor. Appl. Surf. Sci. 2020, 509, 144773. [Google Scholar] [CrossRef]
  434. Shen, D.; Li, X.; Ma, C.; Zhou, Y.; Sun, L.; Yin, S.; Huo, P.; Wang, H. Synthesized Z-Scheme Photocatalyst ZnO/g-C3N4 for Enhanced Photocatalytic Reduction of CO2. New J. Chem. 2020, 44, 16390–16399. [Google Scholar] [CrossRef]
  435. Liang, M.; Borjigin, T.; Zhang, Y.; Liu, B.; Liu, H.; Guo, H. Controlled Assemble of Hollow Heterostructured g-C3N4@ CeO2 with Rich Oxygen Vacancies for Enhanced Photocatalytic CO2 Reduction. Appl. Catal. B Environ. 2019, 243, 566–575. [Google Scholar] [CrossRef]
  436. Tang, J.; Guo, R.; Zhou, W.; Huang, C.; Pan, W. Ball-Flower like NiO/g-C3N4 Heterojunction for Efficient Visible Light Photocatalytic CO2 Reduction. Appl. Catal. B Environ. 2018, 237, 802–810. [Google Scholar] [CrossRef]
  437. Sun, Z.; Fang, W.; Zhao, L.; Chen, H.; He, X.; Li, W.; Tian, P.; Huang, Z. g-C3N4 Foam/Cu2O QDs with Excellent CO2 Adsorption and Synergistic Catalytic Effect for Photocatalytic CO2 Reduction. Environ. Int. 2019, 130, 104898. [Google Scholar] [CrossRef]
  438. Truc, N.T.T.; Pham, T.-D.; Nguyen, M.V.; Van Thuan, D.; Thao, P.; Trang, H.T.; Tran, D.T.; Minh, D.N.; Hanh, N.T.; Ngoc, H.M. Advanced NiMoO4/g-C3N4 Z-Scheme Heterojunction Photocatalyst for Efficient Conversion of CO2 to Valuable Products. J. Alloys Compd. 2020, 842, 155860. [Google Scholar] [CrossRef]
  439. Li, M.; Zhang, L.; Wu, M.; Du, Y.; Fan, X.; Wang, M.; Zhang, L.; Kong, Q.; Shi, J. Mesostructured CeO2/g-C3N4 Nanocomposites: Remarkably Enhanced Photocatalytic Activity for CO2 Reduction by Mutual Component Activations. Nano Energy 2016, 19, 145–155. [Google Scholar] [CrossRef]
  440. Li, X.; Jiang, H.; Ma, C.; Zhu, Z.; Song, X.; Wang, H.; Huo, P.; Li, X. Local Surface Plasma Resonance Effect Enhanced Z-Scheme ZnO/Au/g-C3N4 Film Photocatalyst for Reduction of CO2 to CO. Appl. Catal. B Environ. 2021, 283, 119638. [Google Scholar] [CrossRef]
  441. He, Y.; Wang, Y.; Zhang, L.; Teng, B.; Fan, M. High-Efficiency Conversion of CO2 to Fuel over ZnO/g-C3N4 Photocatalyst. Appl. Catal. B Environ. 2015, 168, 1–8. [Google Scholar] [CrossRef]
  442. Reli, M.; Huo, P.; Šihor, M.; Ambrozova, N.; Troppová, I.; Matejova, L.; Lang, J.; Svoboda, L.; Kuśtrowski, P.; Ritz, M. Novel TiO2/C3N4 Photocatalysts for Photocatalytic Reduction of CO2 and for Photocatalytic Decomposition of N2O. J. Phys. Chem. A 2016, 120, 8564–8573. [Google Scholar] [CrossRef]
  443. Tang, Z.; Wang, C.; He, W.; Wei, Y.; Zhao, Z.; Liu, J. The Z-Scheme g-C3N4/3DOM-WO3 Photocatalysts with Enhanced Activity for CO2 Photoreduction into CO. Chin. Chem. Lett. 2021, in press. [CrossRef]
  444. Guo, H.; Wan, S.; Wang, Y.; Ma, W.; Zhong, Q.; Ding, J. Enhanced Photocatalytic CO2 Reduction over Direct Z-Scheme NiTiO3/g-C3N4 Nanocomposite Promoted by Efficient Interfacial Charge Transfer. Chem. Eng. J. 2021, 412, 128646. [Google Scholar] [CrossRef]
  445. Carmen, Z.; Daniela, S. Textile Organic Dyes-Characteristics, Polluting Effects and Separation/Elimination Procedures from Industrial Effluents-a Critical Overview; IntechOpen: Rijekta, Croatia, 2012. [Google Scholar]
  446. Clarke, E.A.; Anliker, R. Organic Dyes and Pigments. In Anthropogenic Compounds; Springer: Berlin/Heidelberg, Germany, 1980. [Google Scholar]
  447. Danish, M.S.S.; Bhattacharya, A.; Stepanova, D.; Mikhaylov, A.; Grilli, M.L.; Khosravy, M.; Senjyu, T. A Systematic Review of Metal Oxide Applications for Energy and Environmental Sustainability. Metals. 2020, 10, 1604. [Google Scholar] [CrossRef]
  448. Devi, K.R.S.; Mathew, S.; Rajan, R.; Georgekutty, J.; Kasinathan, K.; Pinheiro, D.; Sugunan, S. Biogenic Synthesis of g-C3N4/Bi2O3 Heterojunction with Enhanced Photocatalytic Activity and Statistical Optimization of Reaction Parameters. Appl. Surf. Sci. 2019, 494, 465–476. [Google Scholar] [CrossRef]
  449. Lan, Y.; Li, Z.; Li, D.; Xie, W.; Yan, G.; Guo, S. Visible-Light Responsive Z-Scheme Bi@ β-Bi2O3/g-C3N4 Heterojunction for Efficient Photocatalytic Degradation of 2, 3-Dihydroxynaphthalene. Chem. Eng. J. 2020, 392, 123686. [Google Scholar] [CrossRef]
  450. Khan, A.A.; Tahir, M.; Bafaqeer, A. Constructing a STable 2D Layered Ti3C2 MXene Cocatalyst-Assisted TiO2/g-C3N4/Ti3C2 Heterojunction for Tailoring Photocatalytic Bireforming of Methane under Visible Light. Energy Fuels 2020, 34, 9810–9828. [Google Scholar]
  451. Ma, J.; Tan, X.; Yu, T.; Li, X. Fabrication of g-C3N4/TiO2 Hierarchical Spheres with Reactive {001} TiO2 Crystal Facets and Its Visible-Light Photocatalytic Activity. Int. J. Hydrogen Energy 2016, 41, 3877–3887. [Google Scholar] [CrossRef]
  452. Rasheed, H.U.; Lv, X.; Wei, W.; Yaseen, W.; Ullah, N.; Xie, J.; Zhu, W. Synthesis and Studies of ZnO Doped with g-C3N4 Nanocomposites for the Degradation of Tetracycline Hydrochloride under the Visible Light Irradiation. J. Environ. Chem. Eng. 2019, 7, 103152. [Google Scholar] [CrossRef]
  453. Zhu, W.; Sun, F.; Goei, R.; Zhou, Y. Construction of WO3–g-C3N4 Composites as Efficient Photocatalysts for Pharmaceutical Degradation under Visible Light. Catal. Sci. Technol. 2017, 7, 2591–2600. [Google Scholar] [CrossRef]
  454. Lu, N.; Wang, P.; Su, Y.; Yu, H.; Liu, N.; Quan, X. Construction of Z-Scheme g-C3N4/RGO/WO3 with in Situ Photoreduced Graphene Oxide as Electron Mediator for Efficient Photocatalytic Degradation of Ciprofloxacin. Chemosphere 2019, 215, 444–453. [Google Scholar] [CrossRef]
  455. Truong, H.B.; Huy, B.T.; Ray, S.K.; Lee, Y.-I.; Cho, J.; Hur, J. H2O2-Assisted Photocatalysis for Removal of Natural Organic Matter Using Nanosheet C3N4-WO3 Composite under Visible Light and the Hybrid System with Ultrafiltration. Chem. Eng. J. 2020, 399, 125733. [Google Scholar] [CrossRef]
  456. Shafawi, A.N.; Mahmud, R.A.; Ali, K.A.; Putri, L.K.; Rosli, N.I.M.; Mohamed, A.R. Bi2O3 Particles Decorated on Porous g-C3N4 Sheets: Enhanced Photocatalytic Activity through a Direct Z-Scheme Mechanism for Degradation of Reactive Black 5 under UV–Vis Light. J. Photochem. Photobiol. A Chem. 2020, 389, 112289. [Google Scholar] [CrossRef]
  457. Ma, R.; Zhang, S.; Li, L.; Gu, P.; Wen, T.; Khan, A.; Li, S.; Li, B.; Wang, S.; Wang, X. Enhanced Visible-Light-Induced Photoactivity of Type-II CeO2/g-C3N4 Nanosheet toward Organic Pollutants Degradation. ACS Sustain. Chem. Eng. 2019, 7, 9699–9708. [Google Scholar]
  458. Barathi, D.; Rajalakshmi, N.; Ranjith, R.; Sangeetha, R.; Meyvel, S. Controllable Synthesis of CeO2/g-C3N4 Hybrid Catalysts and Its Structural, Optical and Visible Light Photocatalytic Activity. Diam. Relat. Mater. 2021, 111, 108161. [Google Scholar] [CrossRef]
  459. Cui, S.; Xie, B.; Li, R.; Pei, J.; Tian, Y.; Zhang, J.; Xing, X. g-C3N4/CeO2 Binary Composite Prepared and Its Application in Automobile Exhaust Degradation. Materials. 2020, 13, 1274. [Google Scholar] [CrossRef] [Green Version]
  460. Tan, L.; Xu, J.; Zhang, X.; Hang, Z.; Jia, Y.; Wang, S. Synthesis of g-C3N4/CeO2 Nanocomposites with Improved Catalytic Activity on the Thermal Decomposition of Ammonium Perchlorate. Appl. Surf. Sci. 2015, 356, 447–453. [Google Scholar] [CrossRef]
  461. Chen, G.; Bian, S.; Guo, C.-Y.; Wu, X. Insight into the Z-Scheme Heterostructure WO3/g-C3N4 for Enhanced Photocatalytic Degradation of Methyl Orange. Mater. Lett. 2019, 236, 596–599. [Google Scholar] [CrossRef]
  462. Guan, R.; Li, J.; Zhang, J.; Zhao, Z.; Wang, D.; Zhai, H.; Sun, D. Photocatalytic Performance and Mechanistic Research of ZnO/g-C3N4 on Degradation of Methyl Orange. ACS omega 2019, 4, 20742–20747. [Google Scholar] [CrossRef] [Green Version]
  463. Si, Y.; Zhang, X.; Liang, T.; Xu, X.; Qiu, L.; Li, P.; Duo, S. Facile In-Situ Synthesis of 2D/3D g-C3N4/Cu2O Heterojunction for High-Performance Photocatalytic Dye Degradation. Mater. Res. Express 2020, 7, 15524. [Google Scholar] [CrossRef]
  464. Tian, Q.; Shen, X.; Wang, Z.; Zhu, B.; Osotsi, M.I.; Xie, X.; Jin, Y.; Chen, Z.; Zhang, L. Growth of Cu2O Spherical Superstructures on g-C3N4 as Efficient Visible-Light-Driven p–n Heterojunction Photocatalysts for Degrading Various Organic Pollutants. J. Nanosci. Nanotechnol. 2018, 18, 7355–7363. [Google Scholar] [CrossRef]
  465. Rabé, K.; Liu, L.; Nahyoon, N.A.; Zhang, Y.; Idris, A.M. Enhanced Rhodamine B and Coking Wastewater Degradation and Simultaneous Electricity Generation via Anodic g-C3N4/Fe0 (1%)/TiO2 and Cathodic WO3 in Photocatalytic Fuel Cell System under Visible Light Irradiation. Electrochim. Acta 2019, 298, 430–439. [Google Scholar] [CrossRef]
  466. Yu, T.; Liu, L.; Yang, F. Heterojunction between Anodic TiO2/g-C3N4 and Cathodic WO3/W Nano-Catalysts for Coupled Pollutant Removal in a Self-Biased System. Chin. J. Catal. 2017, 38, 270–277. [Google Scholar] [CrossRef]
  467. Tripathi, A.; Narayanan, S. Impact of TiO2 and TiO2/g-C3N4 Nanocomposite to Treat Industrial Wastewater. Environ. Nanotechnol. Monit. Manag. 2018, 10, 280–291. [Google Scholar] [CrossRef]
  468. Pawar, R.C.; Son, Y.; Kim, J.; Ahn, S.H.; Lee, C.S. Integration of ZnO with g-C3N4 Structures in Core–Shell Approach via Sintering Process for Rapid Detoxification of Water under Visible Irradiation. Curr. Appl. Phys. 2016, 16, 101–108. [Google Scholar] [CrossRef]
  469. Zhang, S.; Su, C.; Ren, H.; Li, M.; Zhu, L.; Ge, S.; Wang, M.; Zhang, Z.; Li, L.; Cao, X. In-Situ Fabrication of g-C3N4/ZnO Nanocomposites for Photocatalytic Degradation of Methylene Blue: Synthesis Procedure Does Matter. Nanomaterials 2019, 9, 215. [Google Scholar] [CrossRef] [Green Version]
  470. Li, X.; Li, M.; Yang, J.; Li, X.; Hu, T.; Wang, J.; Sui, Y.; Wu, X.; Kong, L. Synergistic Effect of Efficient Adsorption g-C3N4/ZnO Composite for Photocatalytic Property. J. Phys. Chem. Solids 2014, 75, 441–446. [Google Scholar] [CrossRef]
  471. Kumaresan, N.; Sinthiya, M.M.A.; Sarathbavan, M.; Ramamurthi, K.; Sethuraman, K.; Babu, R.R. Synergetic Effect of g-C3N4/ZnO Binary Nanocomposites Heterojunction on Improving Charge Carrier Separation through 2D/1D Nanostructures for Effective Photocatalytic Activity under the Sunlight Irradiation. Sep. Purif. Technol. 2020, 244, 116356. [Google Scholar] [CrossRef]
  472. Liu, L.; Luo, X.; Li, Y.; Xu, F.; Gao, Z.; Zhang, X.; Song, Y.; Xu, H.; Li, H. Facile Synthesis of Few-Layer g-C3N4/ZnO Composite Photocatalyst for Enhancing Visible Light Photocatalytic Performance of Pollutants Removal. Colloids Surfaces A Physicochem. Eng. Asp. 2018, 537, 516–523. [Google Scholar] [CrossRef]
  473. Chen, Q.; Hou, H.; Zhang, D.; Hu, S.; Min, T.; Liu, B.; Yang, C.; Pu, W.; Hu, J.; Yang, J. Enhanced Visible-Light Driven Photocatalytic Activity of Hybrid ZnO/g-C3N4 by High Performance Ball Milling. J. Photochem. Photobiol. A Chem. 2018, 350, 1–9. [Google Scholar] [CrossRef]
  474. Tang, Q.; Meng, X.; Wang, Z.; Zhou, J.; Tang, H. One-Step Electrospinning Synthesis of TiO2/g-C3N4 Nanofibers with Enhanced Photocatalytic Properties. Appl. Surf. Sci. 2018, 430, 253–262. [Google Scholar] [CrossRef]
  475. Liu, H.; Zhang, Z.-G.; He, H.-W.; Wang, X.-X.; Zhang, J.; Zhang, Q.-Q.; Tong, Y.-F.; Liu, H.-L.; Ramakrishna, S.; Yan, S.-Y. One-Step Synthesis Heterostructured g-C3N4/TiO2 Composite for Rapid Degradation of Pollutants in Utilizing Visible Light. Nanomaterials 2018, 8, 842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  476. Li, X.; Xiong, J.; Xu, Y.; Feng, Z.; Huang, J. Defect-Assisted Surface Modification Enhances the Visible Light Photocatalytic Performance of g-C3N4@ C-TiO2 Direct Z-Scheme Heterojunctions. Chin. J. Catal. 2019, 40, 424–433. [Google Scholar] [CrossRef]
  477. Xia, Y.; Xu, L.; Peng, J.; Han, J.; Guo, S.; Zhang, L.; Han, Z.; Komarneni, S. TiO2@ g-C3N4 Core/Shell Spheres with Uniform Mesoporous Structures for High Performance Visible-Light Photocatalytic Application. Ceram. Int. 2019, 45, 18844–18851. [Google Scholar] [CrossRef]
  478. Wang, X.; Yang, W.; Li, F.; Xue, Y.; Liu, R.; Hao, Y. In Situ Microwave-Assisted Synthesis of Porous N-TiO2/g-C3N4 Heterojunctions with Enhanced Visible-Light Photocatalytic Properties. Ind. Eng. Chem. Res. 2013, 52, 17140–17150. [Google Scholar] [CrossRef]
  479. Song, G.; Chu, Z.; Jin, W.; Sun, H. Enhanced Performance of g-C3N4/TiO2 Photocatalysts for Degradation of Organic Pollutants under Visible Light. Chin. J. Chem. Eng. 2015, 23, 1326–1334. [Google Scholar] [CrossRef]
  480. Ma, L.; Wang, G.; Jiang, C.; Bao, H.; Xu, Q. Synthesis of Core-Shell TiO2@ g-C3N4 Hollow Microspheres for Efficient Photocatalytic Degradation of Rhodamine B under Visible Light. Appl. Surf. Sci. 2018, 430, 263–272. [Google Scholar] [CrossRef]
  481. Li, K.; Gao, S.; Wang, Q.; Xu, H.; Wang, Z.; Huang, B.; Dai, Y.; Lu, J. In-Situ-Reduced Synthesis of Ti3+ Self-Doped TiO2/g-C3N4 Heterojunctions with High Photocatalytic Performance under LED Light Irradiation. ACS Appl. Mater. Interfaces 2015, 7, 9023–9030. [Google Scholar] [CrossRef] [PubMed]
  482. Yan, X.; Gao, Q.; Qin, J.; Hui, X.; Ye, Z.; Li, J.; Ma, Z. A Facile Method for Fabricating TiO2/g-C3N4 Hollow Nanotube Heterojunction and Its Visible Light Photocatalytic Performance. Mater. Lett. 2018, 217, 1–4. [Google Scholar] [CrossRef]
  483. Christoforidis, K.C. g-C3N4/Ag3PO4 Based Binary and Ternary Heterojunction for Improved Photocatalytic Removal of Organic Pollutants. Int. J. Environ. Anal. Chem. published on line only. 2021. [Google Scholar] [CrossRef]
  484. Dutta, D.P.; Dagar, D. Efficient Selective Sorption of Cationic Organic Pollutant from Water and Its Photocatalytic Degradation by AlVO4/g-C3N4 Nanocomposite. J. Nanosci. Nanotechnol. 2020, 20, 2179–2194. [Google Scholar] [CrossRef]
  485. Hong, Y.; Jiang, Y.; Li, C.; Fan, W.; Yan, X.; Yan, M.; Shi, W. In-Situ Synthesis of Direct Solid-State Z-Scheme V2O5/g-C3N4 Heterojunctions with Enhanced Visible Light Efficiency in Photocatalytic Degradation of Pollutants. Appl. Catal. B Environ. 2016, 180, 663–673. [Google Scholar] [CrossRef]
  486. Meng, J.; Wang, X.; Liu, Y.; Ren, M.; Zhang, X.; Ding, X.; Guo, Y.; Yang, Y. Acid-Induced Molecule Self-Assembly Synthesis of Z-Scheme WO3/g-C3N4 Heterojunctions for Robust Photocatalysis against Phenolic Pollutants. Chem. Eng. J. 2021, 403, 126354. [Google Scholar] [CrossRef]
  487. Liang, J.; Li, Y.; Zhang, J.; Li, C.; Yang, X.; Chen, X.; Wang, F.; Chen, C. Crystalline Phase Engineering in WO3/g-C3N4 Composites for Improved Photocatalytic Performance under Visible Light. Mater. Res. Express 2020, 7, 65503. [Google Scholar] [CrossRef]
  488. He, R.; Zhou, J.; Fu, H.; Zhang, S.; Jiang, C. Room-Temperature in Situ Fabrication of Bi2O3/g-C3N4 Direct Z-Scheme Photocatalyst with Enhanced Photocatalytic Activity. Appl. Surf. Sci. 2018, 430, 273–282. [Google Scholar] [CrossRef]
  489. Chen, D.; Wu, S.; Fang, J.; Lu, S.; Zhou, G.; Feng, W.; Yang, F.; Chen, Y.; Fang, Z. A Nanosheet-like α-Bi2O3/g-C3N4 Heterostructure Modified by Plasmonic Metallic Bi and Oxygen Vacancies with High Photodegradation Activity of Organic Pollutants. Sep. Purif. Technol. 2018, 193, 232–241. [Google Scholar] [CrossRef]
  490. Fan, G.; Ma, Z.; Li, X.; Deng, L. Coupling of Bi2O3 Nanoparticles with g-C3N4 for Enhanced Photocatalytic Degradation of Methylene Blue. Ceram. Int. 2021, 47, 5758–5766. [Google Scholar] [CrossRef]
  491. Ghosh, U.; Pal, A. Fabrication of a Novel Bi2O3 Nanoparticle Impregnated Nitrogen Vacant 2D g-C3N4 Nanosheet Z Scheme Photocatalyst for Improved Degradation of Methylene Blue Dye under LED Light Illumination. Appl. Surf. Sci. 2020, 507, 144965. [Google Scholar] [CrossRef]
  492. She, X.; Xu, H.; Wang, H.; Xia, J.; Song, Y.; Yan, J.; Xu, Y.; Zhang, Q.; Du, D.; Li, H. Controllable Synthesis of CeO2/g-C3N4 Composites and Their Applications in the Environment. Dalt. Trans. 2015, 44, 7021–7031. [Google Scholar] [CrossRef]
  493. Ghalkhani, M.; Charkhan, H.; Sabbaghan, M. Synthesis and Application of a Powerful Heterogeneous Photo-Fenton Catalyst Based on RGO/g-C3N4/Fe3O4/TiO2 Nano-Composite for the Removal of Sewage Contaminants. J. Electrochem. Soc. 2020, 167, 67515. [Google Scholar] [CrossRef]
  494. Vattikuti, S.V.P.; Reddy, P.A.K.; Shim, J.; Byon, C. Visible-Light-Driven Photocatalytic Activity of SnO2–ZnO Quantum Dots Anchored on g-C3N4 Nanosheets for Photocatalytic Pollutant Degradation and H2 Production. ACS omega 2018, 3, 7587–7602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  495. Danish, M.; Muneer, M. Novel ZnSQDs-SnO2/g-C3N4 Nanocomposite with Enhanced Photocatalytic Performance for the Degradation of Different Organic Pollutants in Aqueous Suspension under Visible Light. J. Phys. Chem. Solids 2021, 149, 109785. [Google Scholar] [CrossRef]
  496. Zhang, Z.; Wu, J.; Liu, D. Co3O4/g-C3N4 Hybrids for Gas-Phase Hg0 Removal at Low Temperature. Processes 2019, 7, 279. [Google Scholar] [CrossRef] [Green Version]
  497. Zhao, X.; Lu, Z.; Ji, R.; Zhang, M.; Yi, C.; Yan, Y. Biomass Carbon Modified Z-Scheme g-C3N4/Co3O4 Heterojunction with Enhanced Visible-Light Photocatalytic Activity. Catal. Commun. 2018, 112, 49–52. [Google Scholar] [CrossRef]
  498. Yuan, X.; Duan, S.; Wu, G.; Sun, L.; Cao, G.; Li, D.; Xu, H.; Li, Q.; Xia, D. Enhanced Catalytic Ozonation Performance of Highly Stabilized Mesoporous ZnO Doped g-C3N4 Composite for Efficient Water Decontamination. Appl. Catal. A Gen. 2018, 551, 129–138. [Google Scholar] [CrossRef]
  499. Zuo, S.; Chen, Y.; Liu, W.; Yao, C.; Li, X.; Li, Z.; Ni, C.; Liu, X. A Facile and Novel Construction of Attapulgite/Cu2O/Cu/g-C3N4 with Enhanced Photocatalytic Activity for Antibiotic Degradation. Ceram. Int. 2017, 43, 3324–3329. [Google Scholar] [CrossRef]
  500. Raha, S.; Ahmaruzzaman, M. Enhanced Performance of a Novel Superparamagnetic g-C3N4/NiO/ZnO/Fe3O4 Nanohybrid Photocatalyst for Removal of Esomeprazole: Effects of Reaction Parameters, Co-Existing Substances and Water Matrices. Chem. Eng. J. 2020, 395, 124969. [Google Scholar] [CrossRef]
  501. Huo, R.; Yang, X.-L.; Yang, J.-Y.; Yang, S.-Y.; Xu, Y.-H. Self-Assembly Synthesis of BiVO4/Polydopamine/g-C3N4 with Enhanced Visible Light Photocatalytic Performance. Mater. Res. Bull. 2018, 98, 225–230. [Google Scholar] [CrossRef]
  502. Ding, J.; Lu, S.; Shen, L.; Yan, R.; Zhang, Y.; Zhang, H. Enhanced Photocatalytic Reduction for the Dechlorination of 2-Chlorodibenzo-p-Dioxin by High-Performance g-C3N4/NiO Heterojunction Composites under Ultraviolet-Visible Light Illumination. J. Hazard. Mater. 2020, 384, 121255. [Google Scholar] [CrossRef]
  503. Zhao, W.; Yang, X.; Liu, C.; Qian, X.; Wen, Y.; Yang, Q.; Sun, T.; Chang, W.; Liu, X.; Chen, Z. Facile Construction of All-Solid-State Z-Scheme g-C3N4/TiO2 Thin Film for the Efficient Visible-Light Degradation of Organic Pollutant. Nanomaterials 2020, 10, 600. [Google Scholar] [CrossRef] [Green Version]
  504. Huang, J.; Li, D.; Li, R.; Chen, P.; Zhang, Q.; Liu, H.; Lv, W.; Liu, G.; Feng, Y. One-Step Synthesis of Phosphorus/Oxygen Co-Doped g-C3N4/Anatase TiO2 Z-Scheme Photocatalyst for Significantly Enhanced Visible-Light Photocatalysis Degradation of Enrofloxacin. J. Hazard. Mater. 2020, 386, 121634. [Google Scholar] [CrossRef]
  505. Hao, J.; Zhang, S.; Ren, F.; Wang, Z.; Lei, J.; Wang, X.; Cheng, T.; Li, L. Synthesis of TiO2@ g-C3N4 Core-Shell Nanorod Arrays with Z-Scheme Enhanced Photocatalytic Activity under Visible Light. J. Colloid Interface Sci. 2017, 508, 419–425. [Google Scholar] [CrossRef]
  506. Karpuraranjith, M.; Chen, Y.; Rajaboopathi, S.; Ramadoss, M.; Srinivas, K.; Yang, D.; Wang, B. Three-Dimensional Porous MoS2 Nanobox Embedded g-C3N4@ TiO2 Architecture for Highly Efficient Photocatalytic Degradation of Organic Pollutant. J. Colloid Interface Sci. 2022, 605, 613–623. [Google Scholar] [CrossRef] [PubMed]
  507. Zhou, Q.; Zhao, D.; Sun, Y.; Sheng, X.; Zhao, J.; Guo, J.; Zhou, B. G–C3N4–and Polyaniline-Co-Modified TiO2 Nanotube Arrays for Significantly Enhanced Photocatalytic Degradation of Tetrabromobisphenol A under Visible Light. Chemosphere 2020, 252, 126468. [Google Scholar] [CrossRef] [PubMed]
  508. Wang, Y.; Rao, L.; Wang, P.; Shi, Z.; Zhang, L. Photocatalytic Activity of N-TiO2/O-Doped N Vacancy g-C3N4 and the Intermediates Toxicity Evaluation under Tetracycline Hydrochloride and Cr (VI) Coexistence Environment. Appl. Catal. B Environ. 2020, 262, 118308. [Google Scholar] [CrossRef]
  509. Wang, W.; Fang, J.; Shao, S.; Lai, M.; Lu, C. Compact and Uniform TiO2@ g-C3N4 Core-Shell Quantum Heterojunction for Photocatalytic Degradation of Tetracycline Antibiotics. Appl. Catal. B Environ. 2017, 217, 57–64. [Google Scholar] [CrossRef]
  510. Du, X.; Bai, X.; Xu, L.; Yang, L.; Jin, P. Visible-Light Activation of Persulfate by TiO2/g-C3N4 Photocatalyst toward Efficient Degradation of Micropollutants. Chem. Eng. J. 2020, 384, 123245. [Google Scholar] [CrossRef]
  511. Yang, Z.; Yan, J.; Lian, J.; Xu, H.; She, X.; Li, H. g-C3N4/TiO2 Nanocomposites for Degradation of Ciprofloxacin under Visible Light Irradiation. ChemistrySelect 2016, 1, 5679–5685. [Google Scholar] [CrossRef]
  512. Ngullie, R.C.; Alaswad, S.O.; Bhuvaneswari, K.; Shanmugam, P.; Pazhanivel, T.; Arunachalam, P. Synthesis and Characterization of Efficient ZnO/g-C3N4 Nanocomposites Photocatalyst for Photocatalytic Degradation of Methylene Blue. Coatings 2020, 10, 500. [Google Scholar] [CrossRef]
  513. Li, N.; Tian, Y.; Zhao, J.; Zhang, J.; Zuo, W.; Kong, L.; Cui, H. Z-Scheme 2D/3D g-C3N4@ ZnO with Enhanced Photocatalytic Activity for Cephalexin Oxidation under Solar Light. Chem. Eng. J. 2018, 352, 412–422. [Google Scholar] [CrossRef]
  514. Prabhu, S.; Pudukudy, M.; Harish, S.; Navaneethan, M.; Sohila, S.; Murugesan, K.; Ramesh, R. Facile Construction of Djembe-like ZnO and Its Composite with g-C3N4 as a Visible-Light-Driven Heterojunction Photocatalyst for the Degradation of Organic Dyes. Mater. Sci. Semicond. Process. 2020, 106, 104754. [Google Scholar] [CrossRef]
  515. Pan, T.; Chen, D.; Xu, W.; Fang, J.; Wu, S.; Liu, Z.; Wu, K.; Fang, Z. Anionic Polyacrylamide-Assisted Construction of Thin 2D-2D WO3/g-C3N4 Step-Scheme Heterojunction for Enhanced Tetracycline Degradation under Visible Light Irradiation. J. Hazard. Mater. 2020, 393, 122366. [Google Scholar] [CrossRef] [PubMed]
  516. Priya, A.; Senthil, R.A.; Selvi, A.; Arunachalam, P.; Kumar, C.K.S.; Madhavan, J.; Boddula, R.; Pothu, R.; Al-Mayouf, A.M. A Study of Photocatalytic and Photoelectrochemical Activity of As-Synthesized WO3/g-C3N4 Composite Photocatalysts for AO7 Degradation. Mater. Sci. Energy Technol. 2020, 3, 43–50. [Google Scholar] [CrossRef]
  517. Navarro-Aguilar, A.I.; Obregón, S.; Sanchez-Martinez, D.; Hernández-Uresti, D.B. An Efficient and Stable WO3/g-C3N4 Photocatalyst for Ciprofloxacin and Orange G Degradation. J. Photochem. Photobiol. A Chem. 2019, 384, 112010. [Google Scholar] [CrossRef]
  518. Manikandan, V.S.; Harish, S.; Archana, J.; Navaneethan, M. Fabrication of Novel Hybrid Z-Scheme WO3@g-C3N4@ MWCNT Nanostructure for Photocatalytic Degradation of Tetracycline and the Evaluation of Antimicrobial Activity. Chemosphere 2021, 287, 132050. [Google Scholar] [CrossRef] [PubMed]
  519. Zhou, S.; Wang, Y.; Zhou, K.; Ba, D.; Ao, Y.; Wang, P. In-Situ Construction of Z-Scheme g-C3N4/WO3 Composite with Enhanced Visible-Light Responsive Performance for Nitenpyram Degradation. Chin. Chem. Lett. 2021, 32, 2179–2182. [Google Scholar] [CrossRef]
  520. Zhang, L.; Wang, G.; Xiong, Z.; Tang, H.; Jiang, C. Fabrication of Flower-like Direct Z-Scheme β-Bi2O3/g-C3N4 Photocatalyst with Enhanced Visible Light Photoactivity for Rhodamine B Degradation. Appl. Surf. Sci. 2018, 436, 162–171. [Google Scholar] [CrossRef]
  521. Hong, Y.; Li, C.; Yin, B.; Li, D.; Zhang, Z.; Mao, B.; Fan, W.; Gu, W.; Shi, W. Promoting Visible-Light-Induced Photocatalytic Degradation of Tetracycline by an Efficient and Stable Beta-Bi2O3@ g-C3N4 Core/Shell Nanocomposite. Chem. Eng. J. 2018, 338, 137–146. [Google Scholar] [CrossRef]
  522. Liu, W.; Zhou, J.; Hu, Z. Nano-Sized g-C3N4 Thin Layer@ CeO2 Sphere Core-Shell Photocatalyst Combined with H2O2 to Degrade Doxycycline in Water under Visible Light Irradiation. Sep. Purif. Technol. 2019, 227, 115665. [Google Scholar] [CrossRef]
  523. Liu, W.; Zhou, J.; Yao, J. Shuttle-like CeO2/g-C3N4 Composite Combined with Persulfate for the Enhanced Photocatalytic Degradation of Norfloxacin under Visible Light. Ecotoxicol. Environ. Saf. 2020, 190, 110062. [Google Scholar] [CrossRef] [PubMed]
  524. Huang, Z.; Li, L.; Li, Z.; Li, H.; Wu, J. Synthesis of Novel Kaolin-Supported g-C3N4/CeO2 Composites with Enhanced Photocatalytic Removal of Ciprofloxacin. Materials. 2020, 13, 3811. [Google Scholar] [CrossRef] [PubMed]
  525. Wang, S.; Long, J.; Jiang, T.; Shao, L.; Li, D.; Xie, X.; Xu, F. Magnetic Fe3O4/CeO2/g-C3N4 Composites with a Visible-Light Response as a High Efficiency Fenton Photocatalyst to Synergistically Degrade Tetracycline. Sep. Purif. Technol. 2021, 278, 119609. [Google Scholar] [CrossRef]
  526. Han, C.; Ge, L.; Chen, C.; Li, Y.; Xiao, X.; Zhang, Y.; Guo, L. Novel Visible Light Induced Co3O4-g-C3N4 Heterojunction Photocatalysts for Efficient Degradation of Methyl Orange. Appl. Catal. B Environ. 2014, 147, 546–553. [Google Scholar] [CrossRef]
  527. Chen, H.-Y.; Qiu, L.-G.; Xiao, J.-D.; Ye, S.; Jiang, X.; Yuan, Y.-P. Inorganic–Organic Hybrid NiO–g-C3N4 Photocatalyst for Efficient Methylene Blue Degradation Using Visible Light. RSC Adv. 2014, 4, 22491–22496. [Google Scholar] [CrossRef]
  528. Balasubramanian, P.; Annalakshmi, M.; Chen, S.-M.; Chen, T.-W. Sonochemical Synthesis of Molybdenum Oxide (MoO3) Microspheres Anchored Graphitic Carbon Nitride (g-C3N4) Ultrathin Sheets for Enhanced Electrochemical Sensing of Furazolidone. Ultrason. Sonochem. 2019, 50, 96–104. [Google Scholar] [CrossRef] [PubMed]
  529. Adhikari, S.; Kim, D.-H. Heterojunction C3N4/MoO3 Microcomposite for Highly Efficient Photocatalytic Oxidation of Rhodamine B. Appl. Surf. Sci. 2020, 511, 145595. [Google Scholar] [CrossRef]
  530. Xue, S.; Wu, C.; Pu, S.; Hou, Y.; Tong, T.; Yang, G.; Qin, Z.; Wang, Z.; Bao, J. Direct Z-Scheme Charge Transfer in Heterostructured MoO3/g-C3N4 Photocatalysts and the Generation of Active Radicals in Photocatalytic Dye Degradations. Environ. Pollut. 2019, 250, 338–345. [Google Scholar] [CrossRef]
  531. Liu, L.; Huang, J.; Yu, H.; Wan, J.; Liu, L.; Yi, K.; Zhang, W.; Zhang, C. Construction of MoO3 Nanopaticles/g-C3N4 Nanosheets 0D/2D Heterojuntion Photocatalysts for Enhanced Photocatalytic Degradation of Antibiotic Pollutant. Chemosphere 2021, 282, 131049. [Google Scholar] [CrossRef] [PubMed]
  532. Adorna Jr, J.; Annadurai, T.; Bui, T.A.N.; Tran, H.L.; Lin, L.-Y.; Doong, R.-A. Indirect Z-Scheme Nitrogen-Doped Carbon Dot Decorated Bi2MoO6/g-C3N4 Photocatalyst for Enhanced Visible-Light-Driven Degradation of Ciprofloxacin. Chem. Eng. J. 2021, 422, 130103. [Google Scholar]
  533. Van Pham, V.; Mai, D.-Q.; Bui, D.-P.; Van Man, T.; Zhu, B.; Zhang, L.; Sangkaworn, J.; Tantirungrotechai, J.; Reutrakul, V.; Cao, T.M. Emerging 2D/0D g-C3N4/SnO2 S-Scheme Photocatalyst: New Generation Architectural Structure of Heterojunctions toward Visible-Light-Driven NO Degradation. Environ. Pollut. 2021, 286, 117510. [Google Scholar] [CrossRef]
  534. Shi, Y.; Yang, Z.; Liu, Y.; Yu, J.; Wang, F.; Tong, J.; Su, B.; Wang, Q. Fabricating a g-C3N4/CuO x Heterostructure with Tunable Valence Transition for Enhanced Photocatalytic Activity. RSC Adv. 2016, 6, 39774–39783. [Google Scholar] [CrossRef]
  535. Yuan, X.; Zhou, C.; Jing, Q.; Tang, Q.; Mu, Y.; Du, A. Facile Synthesis of g-C3N4 Nanosheets/ZnO Nanocomposites with Enhanced Photocatalytic Activity in Reduction of Aqueous Chromium (VI) under Visible Light. Nanomaterials 2016, 6, 173. [Google Scholar] [CrossRef] [Green Version]
  536. Sundaram, I.M.; Kalimuthu, S.; Ponniah, G. Highly Active ZnO Modified g-C3N4 Nanocomposite for Dye Degradation under UV and Visible Light with Enhanced Stability and Antimicrobial Activity. Compos. Commun. 2017, 5, 64–71. [Google Scholar] [CrossRef]
  537. Vattikuti, S.V.P.; Byon, C. Hydrothermally Synthesized Ternary Heterostructured MoS2/Al2O3/g-C3N4 Photocatalyst. Mater. Res. Bull. 2017, 96, 233–245. [Google Scholar] [CrossRef]
  538. Wang, J.; Xia, Y.; Zhao, H.; Wang, G.; Xiang, L.; Xu, J.; Komarneni, S. Oxygen Defects-Mediated Z-Scheme Charge Separation in g-C3N4/ZnO Photocatalysts for Enhanced Visible-Light Degradation of 4-Chlorophenol and Hydrogen Evolution. Appl. Catal. B Environ. 2017, 206, 406–416. [Google Scholar] [CrossRef] [Green Version]
  539. Zhou, Y.; Zeng, F.; Sun, C.; Wu, J.; Xie, Y.; Zhang, F.; Rao, S.; Wang, F.; Zhang, J.; Zhao, J. Gd2O3 Nanoparticles Modified g-C3N4 with Enhanced Photocatalysis Activity for the Degradation of Organic Pollutants1. J. Rare Earths 2021, 39, 1353–1361. [Google Scholar] [CrossRef]
  540. Chen, D.; Xie, Z.; Zeng, Y.; Lv, W.; Zhang, Q.; Wang, F.; Liu, G.; Liu, H. Accelerated Photocatalytic Degradation of Quinolone Antibiotics over Z-Scheme MoO3/g-C3N4 Heterostructure by Peroxydisulfate under Visible Light Irradiation: Mechanism; Kinetic; and Products. J. Taiwan Inst. Chem. Eng. 2019, 104, 250–259. [Google Scholar] [CrossRef]
  541. Devi KR, S.; Mathew, S.; Rajan, R.; Georgekutty, J.; Pinheiro, D.; Ananthapadmanabhan, U.; Sundararajan, M. Synthesis and Characterization of CeO2/Bi2O3/GC3N4 Ternary Z-scheme Nanocomposite. Int. J. Appl. Ceram. Technol. 2020, 17, 2346–2356. [Google Scholar] [CrossRef]
  542. Bajiri, M.A.; Hezam, A.; Namratha, K.; Viswanath, R.; Drmosh, Q.A.; Naik, H.S.B.; Byrappa, K. CuO/ZnO/g-C3N4 Heterostructures as Efficient Visible Light-Driven Photocatalysts. J. Environ. Chem. Eng. 2019, 7, 103412. [Google Scholar] [CrossRef]
  543. Raha, S.; Ahmaruzzaman, M. Facile Fabrication of g-C3N4 Supported Fe3O4 Nanoparticles/ZnO Nanorods: A Superlative Visible Light Responsive Architecture for Express Degradation of Pantoprazole. Chem. Eng. J. 2020, 387, 123766. [Google Scholar] [CrossRef]
  544. Raza, A.; Shen, H.; Haidry, A.A.; Cui, S. Hydrothermal Synthesis of Fe3O4/TiO2/g-C3N4: Advanced Photocatalytic Application. Appl. Surf. Sci. 2019, 488, 887–895. [Google Scholar] [CrossRef]
  545. Jiang, D.; Yu, H.; Yu, H. Modified g-C3N4/TiO2 Nanosheets/ZnO Ternary Facet Coupled Heterojunction for Photocatalytic Degradation of p-Toluenesulfonic Acid (p-TSA) under Visible Light. Phys. E Low-dimensional Syst. Nanostruct. 2017, 85, 1–6. [Google Scholar]
  546. Tahir, M.B.; Sagir, M.; Shahzad, K. Removal of Acetylsalicylate and Methyl-Theobromine from Aqueous Environment Using Nano-Photocatalyst WO3-TiO2@g-C3N4 Composite. J. Hazard. Mater. 2019, 363, 205–213. [Google Scholar] [CrossRef] [PubMed]
  547. Yu, B.; Meng, F.; Khan, M.W.; Qin, R.; Liu, X. Synthesis of Hollow TiO2@ g-C3N4/Co3O4 Core-Shell Microspheres for Effective Photooxidation Degradation of Tetracycline and MO. Ceram. Int. 2020, 46, 13133–13143. [Google Scholar] [CrossRef]
  548. Zhang, Y.; Lu, J.; Hoffmann, M.R.; Wang, Q.; Cong, Y.; Wang, Q.; Jin, H. Synthesis of g-C3N4/Bi2O3/TiO2 Composite Nanotubes: Enhanced Activity under Visible Light Irradiation and Improved Photoelectrochemical Activity. RSC Adv. 2015, 5, 48983–48991. [Google Scholar] [CrossRef] [Green Version]
  549. Wang, Y.; Zhang, R.; Zhang, Z.; Cao, J.; Ma, T. Host–Guest Recognition on 2D Graphitic Carbon Nitride for Nanosensing. Adv. Mater. Interfaces 2019, 6, 1901429. [Google Scholar]
  550. Liu, J.-W.; Luo, Y.; Wang, Y.-M.; Duan, L.-Y.; Jiang, J.-H.; Yu, R.-Q. Graphitic Carbon Nitride Nanosheets-Based Ratiometric Fluorescent Probe for Highly Sensitive Detection of H2O2 and Glucose. ACS Appl. Mater. Interfaces 2016, 8, 33439–33445. [Google Scholar] [CrossRef]
  551. Kundu, M.K.; Sadhukhan, M.; Barman, S. Ordered Assemblies of Silver Nanoparticles on Carbon Nitride Sheets and Their Application in the Non-Enzymatic Sensing of Hydrogen Peroxide and Glucose. J. Mater. Chem. B 2015, 3, 1289–1300. [Google Scholar] [CrossRef] [PubMed]
  552. Jiang, T.; Jiang, G.; Huang, Q.; Zhou, H. High-Sensitive Detection of Dopamine Using Graphitic Carbon Nitride by Electrochemical Method. Mater. Res. Bull. 2016, 74, 271–277. [Google Scholar] [CrossRef]
  553. Zuo, F.; Jin, L.; Fu, X.; Zhang, H.; Yuan, R.; Chen, S. An Electrochemiluminescent Sensor for Dopamine Detection Based on a Dual-Molecule Recognition Strategy and Polyaniline Quenching. Sensors Actuators B Chem. 2017, 244, 282–289. [Google Scholar] [CrossRef]
  554. Zou, J.; Wu, S.; Liu, Y.; Sun, Y.; Cao, Y.; Hsu, J.-P.; Wee, A.T.S.; Jiang, J. An Ultra-Sensitive Electrochemical Sensor Based on 2D g-C3N4/CuO Nanocomposites for Dopamine Detection. Carbon N. Y. 2018, 130, 652–663. [Google Scholar] [CrossRef]
  555. Rahman, N.; Yang, J.; Sohail, M.; Khan, R.; Iqbal, A.; Maouche, C.; Khan, A.A.; Husain, M.; Khattak, S.A.; Khan, S.N. Insight into Metallic Oxide Semiconductor (SnO2, ZnO, CuO, α-Fe2O3, WO3)-Carbon Nitride (g-C3N4) Heterojunction for Gas Sensing Application. Sensors Actuators A Phys. 2021, 332, 113128. [Google Scholar] [CrossRef]
  556. Zhou, T.; Zhang, T. Recent Progress of Nanostructured Sensing Materials from 0D to 3D: Overview of Structure–Property-Application Relationship for Gas Sensors. Small Methods 2021, 5, 2100515. [Google Scholar] [CrossRef]
  557. Li, P.-P.; Cao, Y.; Mao, C.-J.; Jin, B.-K.; Zhu, J.-J. TiO2/g-C3N4/CdS Nanocomposite-Based Photoelectrochemical Biosensor for Ultrasensitive Evaluation of T4 Polynucleotide Kinase Activity. Anal. Chem. 2018, 91, 1563–1570. [Google Scholar] [CrossRef] [PubMed]
  558. Li, X.; Zhu, L.; Zhou, Y.; Yin, H.; Ai, S. Enhanced Photoelectrochemical Method for Sensitive Detection of Protein Kinase A Activity Using TiO2/g-C3N4, PAMAM Dendrimer, and Alkaline Phosphatase. Anal. Chem. 2017, 89, 2369–2376. [Google Scholar] [CrossRef]
  559. Zhai, J.; Wang, T.; Wang, C.; Liu, D. UV-Light-Assisted Ethanol Sensing Characteristics of g-C3N4/ZnO Composites at Room Temperature. Appl. Surf. Sci. 2018, 441, 317–323. [Google Scholar] [CrossRef]
  560. Li, X.; Li, Y.; Sun, G.; Luo, N.; Zhang, B.; Zhang, Z. Synthesis of a Flower-like g-C3N4/ZnO Hierarchical Structure with Improved CH4 Sensing Properties. Nanomaterials 2019, 9, 724. [Google Scholar] [CrossRef] [Green Version]
  561. Wang, H.; Bai, J.; Dai, M.; Liu, K.; Liu, Y.; Zhou, L.; Liu, F.; Liu, F.; Gao, Y.; Yan, X. Visible Light Activated Excellent NO2 Sensing Based on 2D/2D ZnO/g-C3N4 Heterojunction Composites. Sens. Actuators B Chem. 2020, 304, 127287. [Google Scholar] [CrossRef]
  562. Zeng, B.; Zhang, L.; Wan, X.; Song, H.; Lv, Y. Fabrication of α-Fe2O3/g-C3N4 Composites for Cataluminescence Sensing of H2S. Sensors Actuators B Chem. 2015, 211, 370–376. [Google Scholar] [CrossRef]
  563. Zhang, Y.; Liu, J.; Chu, X.; Liang, S.; Kong, L. Preparation of g-C3N4–SnO2 Composites for Application as Acetic Acid Sensor. J. Alloys Compd. 2020, 832, 153355. [Google Scholar] [CrossRef]
  564. Malik, R.; Tomer, V.K.; Chaudhary, V.; Dahiya, M.S.; Sharma, A.; Nehra, S.P.; Duhan, S.; Kailasam, K. An Excellent Humidity Sensor Based on In–SnO2 Loaded Mesoporous Graphitic Carbon Nitride. J. Mater. Chem. A 2017, 5, 14134–14143. [Google Scholar] [CrossRef]
  565. Sun, Y.; Jiang, J.; Liu, Y.; Wu, S.; Zou, J. A Facile One-Pot Preparation of Co3O4/g-C3N4 Heterojunctions with Excellent Electrocatalytic Activity for the Detection of Environmental Phenolic Hormones. Appl. Surf. Sci. 2018, 430, 362–370. [Google Scholar] [CrossRef]
  566. Hassannezhad, M.; Hosseini, M.; Ganjali, M.R.; Arvand, M. A Graphitic Carbon Nitride (g-C3N4/Fe3O4) Nanocomposite: An Efficient Electrode Material for the Electrochemical Determination of Tramadol in Human Biological Fluids. Anal. Methods 2019, 11, 2064–2071. [Google Scholar] [CrossRef]
  567. Karthika, A.; Suganthi, A.; Rajarajan, M. An In-Situ Synthesis of Novel V2O5/g-C3N4/PVA Nanocomposite for Enhanced Electrocatalytic Activity toward Sensitive and Selective Sensing of Folic Acid in Natural Samples. Arab. J. Chem. 2020, 13, 3639–3652. [Google Scholar] [CrossRef]
  568. Selvarajan, S.; Suganthi, A.; Rajarajan, M. Fabrication of g-C3N4/NiO Heterostructured Nanocomposite Modified Glassy Carbon Electrode for Quercetin Biosensor. Ultrason. Sonochem. 2018, 41, 651–660. [Google Scholar] [CrossRef]
  569. Wang, L.; Zhao, F.; Han, Q.; Hu, C.; Lv, L.; Chen, N.; Qu, L. Spontaneous Formation of Cu2O–g-C3N4 Core–Shell Nanowires for Photocurrent and Humidity Responses. Nanoscale 2015, 7, 9694–9702. [Google Scholar] [CrossRef] [PubMed]
  570. Liu, Y.; Jiang, J.; Sun, Y.; Wu, S.; Cao, Y.; Gong, W.; Zou, J. NiO and Co3O4 Co-Doped g-C3N4 Nanocomposites with Excellent Photoelectrochemical Properties under Visible Light for Detection of Tetrabromobisphenol-A. RSC Adv. 2017, 7, 36015–36020. [Google Scholar] [CrossRef] [Green Version]
  571. Pang, X.; Cui, C.; Su, M.; Wang, Y.; Wei, Q.; Tan, W. Construction of Self-Powered Cytosensing Device Based on ZnO Nanodisks@ g-C3N4 Quantum Dots and Application in the Detection of CCRF-CEM Cells. Nano Energy 2018, 46, 101–109. [Google Scholar] [CrossRef]
  572. Han, Z.; Luo, M.; Weng, Q.; Chen, L.; Chen, J.; Li, C.; Zhou, Y.; Wang, L. ZnO Flower-Rod/g-C3N4-Gold Nanoparticle-Based Photoelectrochemical Aptasensor for Detection of Carcinoembryonic Antigen. Anal. Bioanal. Chem. 2018, 410, 6529–6538. [Google Scholar] [CrossRef]
  573. Huan, Z.; Chang, J.; Zhou, J. Low-Temperature Fabrication of Macroporous Scaffolds through Foaming and Hydration of Tricalcium Silicate Paste and Their Bioactivity. J. Mater. Sci. 2010, 45, 961–968. [Google Scholar] [CrossRef] [Green Version]
  574. He, X.; Bai, S.; Jiang, J.; Ong, W.-J.; Peng, J.; Xiong, Z.; Liao, G.; Zou, J.; Li, N. Oxygen Vacancy Mediated Step-Scheme Heterojunction of WO2. 9/g-C3N4 for Efficient Electrochemical Sensing of 4-Nitrophenol. Chem. Eng. J. Adv. 2021, 8, 100175. [Google Scholar] [CrossRef]
  575. Wang, H.; Liang, D.; Xu, Y.; Liang, X.; Qiu, X.; Lin, Z. A Highly Efficient Photoelectrochemical Sensor for Detection of Chlorpyrifos Based on 2D/2D β-Bi2O3/g-C3N4 Heterojunctions. Environ. Sci. Nano 2021, 8, 773–783. [Google Scholar]
  576. Pu, Y.; Wu, Y.; Yu, Z.; Lu, L.; Wang, X. Simultaneous Determination of Cd2+ and Pb2+ by an Electrochemical Sensor Based on Fe3O4/Bi2O3/C3N4 Nanocomposites. Talanta Open 2021, 3, 100024. [Google Scholar] [CrossRef]
  577. Bilal, S.; Hassan, M.M.; ur Rehman, M.F.; Nasir, M.; Sami, A.J.; Hayat, A. An Insect Acetylcholinesterase Biosensor Utilizing WO3/g-C3N4 Nanocomposite Modified Pencil Graphite Electrode for Phosmet Detection in Stored Grains. Food Chem. 2021, 346, 128894. [Google Scholar] [CrossRef]
  578. Peng, B.; Zhang, Z.; Tang, L.; Ouyang, X.; Zhu, X.; Chen, L.; Fan, X.; Zhou, Z.; Wang, J. Self-Powered Photoelectrochemical Aptasensor for Oxytetracycline Cathodic Detection Based on a Dual Z-Scheme WO3/g-C3N4/MnO2 Photoanode. Anal. Chem. 2021, 26, 9129–9138. [Google Scholar] [CrossRef]
  579. Mao, L.; Xue, X.; Xu, X.; Wen, W.; Chen, M.-M.; Zhang, X.; Wang, S. Heterostructured CuO-g-C3N4 Nanocomposites as a Highly Efficient Photocathode for Photoelectrochemical Aflatoxin B1 Sensing. Sensors Actuators B Chem. 2021, 329, 129146. [Google Scholar] [CrossRef]
  580. Matsunaga, T.; Tomoda, R.; Nakajima, T.; Wake, H. Photoelectrochemical Sterilization of Microbial Cells by Semiconductor Powders. FEMS Microbiol. Lett. 1985, 29, 211–214. [Google Scholar] [CrossRef]
  581. Song, J.; Wang, X.; Ma, J.; Wang, X.; Wang, J.; Xia, S.; Zhao, J. Removal of Microcystis Aeruginosa and Microcystin-LR Using a Graphitic-C3N4/TiO2 Floating Photocatalyst under Visible Light Irradiation. Chem. Eng. J. 2018, 348, 380–388. [Google Scholar] [CrossRef]
  582. Chen, X.; Wang, Q.; Tian, J.; Liu, Y.; Wang, Y.; Yang, C. A Study on the Photocatalytic Sterilization Performance and Mechanism of Fe-SnO2/g-C3N4 Heterojunction Materials. New J. Chem. 2020, 44, 9456–9465. [Google Scholar] [CrossRef]
  583. Pant, B.; Park, M.; Lee, J.H.; Kim, H.-Y.; Park, S.-J. Novel Magnetically Separable Silver-Iron Oxide Nanoparticles Decorated Graphitic Carbon Nitride Nano-Sheets: A Multifunctional Photocatalyst via One-Step Hydrothermal Process. J. Colloid Interface Sci. 2017, 496, 343–352. [Google Scholar] [CrossRef]
  584. Liu, B.; Wu, Y.; Zhang, J.; Han, X.; Shi, H. Visible-Light-Driven g-C3N4/Cu2O Heterostructures with Efficient Photocatalytic Activities for Tetracycline Degradation and Microbial Inactivation. J. Photochem. Photobiol. A Chem. 2019, 378, 1–8. [Google Scholar] [CrossRef]
  585. Vignesh, S.; Suganthi, S.; Sundar, J.K.; Raj, V. Construction of α-Fe2O3/CeO2 Decorated g-C3N4 Nanosheets for Magnetically Separable Efficient Photocatalytic Performance under Visible Light Exposure and Bacterial Disinfection. Appl. Surf. Sci. 2019, 488, 763–777. [Google Scholar] [CrossRef]
  586. Shanmugam, V.; Jeyaperumal, K.S.; Mariappan, P.; Muppudathi, A.L. Fabrication of Novel g-C3N4 Based MoS 2 and Bi2O3 Nanorod Embedded Ternary Nanocomposites for Superior Photocatalytic Performance and Destruction of Bacteria. New J. Chem. 2020, 44, 13182–13194. [Google Scholar] [CrossRef]
  587. Induja, M.; Sivaprakash, K.; Karthikeyan, S. Facile Green Synthesis and Antimicrobial Performance of Cu2O Nanospheres Decorated g-C3N4 Nanocomposite. Mater. Res. Bull. 2019, 112, 331–335. [Google Scholar] [CrossRef]
  588. Li, C.; Sun, Z.; Zhang, W.; Yu, C.; Zheng, S. Highly Efficient g-C3N4/TiO2/Kaolinite Composite with Novel Three-Dimensional Structure and Enhanced Visible Light Responding Ability towards Ciprofloxacin and S. Aureus. Appl. Catal. B Environ. 2018, 220, 272–282. [Google Scholar] [CrossRef]
  589. Zhang, Q.; Quan, X.; Wang, H.; Chen, S.; Su, Y.; Li, Z. Constructing a Visible-Light-Driven Photocatalytic Membrane by g-C3N4 Quantum Dots and TiO2 Nanotube Array for Enhanced Water Treatment. Sci. Rep. 2017, 7, 1–7. [Google Scholar]
  590. Xu, J.; Li, Y.; Zhou, X.; Li, Y.; Gao, Z.; Song, Y.; Schmuki, P. Graphitic C3N4-Sensitized TiO2 Nanotube Layers: A Visible-Light Activated Efficient Metal-Free Antimicrobial Platform. Chem. Eur. J. 2016, 22, 3947–3951. [Google Scholar] [CrossRef] [Green Version]
  591. Adhikari, S.P.; Awasthi, G.P.; Lee, J.; Park, C.H.; Kim, C.S. Synthesis, Characterization, Organic Compound Degradation Activity and Antimicrobial Performance of g-C3N4 Sheets Customized with Metal Nanoparticles-Decorated TiO2 Nanofibers. RSC Adv. 2016, 6, 55079–55091. [Google Scholar] [CrossRef]
  592. Jiang, G.; Cao, J.; Chen, M.; Zhang, X.; Dong, F. Photocatalytic NO Oxidation on N-Doped TiO2/g-C3N4 Heterojunction: Enhanced Efficiency, Mechanism and Reaction Pathway. Appl. Surf. Sci. 2018, 458, 77–85. [Google Scholar] [CrossRef]
  593. Vo, V.; Thi, X.D.N.; Jin, Y.-S.; Thi, G.L.; Nguyen, T.T.; Duong, T.Q.; Kim, S.-J. SnO2 Nanosheets/g-C3N4 Composite with Improved Lithium Storage Capabilities. Chem. Phys. Lett. 2017, 674, 42–47. [Google Scholar] [CrossRef]
  594. Montazeri, S.M.; Sadrnezhaad, S.K. Kinetics of Sulfur Removal from Tehran Vehicular Gasoline by g-C3N4/SnO2 Nanocomposite. ACS omega 2019, 4, 13180–13188. [Google Scholar] [CrossRef] [Green Version]
  595. Li, X.; Feng, Y.; Li, M.; Li, W.; Wei, H.; Song, D. Smart Hybrids of Zn2GeO4 Nanoparticles and Ultrathin g-C3N4 Layers: Synergistic Lithium Storage and Excellent Electrochemical Performance. Adv. Funct. Mater. 2015, 25, 6858–6866. [Google Scholar] [CrossRef]
  596. Mohamed, H.S.H.; Wu, L.; Li, C.-F.; Hu, Z.-Y.; Liu, J.; Deng, Z.; Chen, L.-H.; Li, Y.; Su, B.-L. In-Situ Growing Mesoporous CuO/O-Doped g-C3N4 Nanospheres for Highly Enhanced Lithium Storage. ACS Appl. Mater. Interfaces 2019, 11, 32957–32968. [Google Scholar] [CrossRef] [PubMed]
  597. Zhang, Y.; Chang, L.; Chang, X.; Chen, H.; Li, Y.; Fan, Y.; Wang, J.; Cui, D.; Xue, C. Combining In-Situ Sedimentation and Carbon-Assisted Synthesis of Co3O4/g-C3N4 Nanocomposites for Improved Supercapacitor Performance. Diam. Relat. Mater. 2021, 111, 108165. [Google Scholar] [CrossRef]
  598. Bai, L.; Huang, H.; Zhang, S.; Hao, L.; Zhang, Z.; Li, H.; Sun, L.; Guo, L.; Huang, H.; Zhang, Y. Photocatalysis-Assisted Co3O4/g-C3N4 p–n Junction All-Solid-State Supercapacitors: A Bridge between Energy Storage and Photocatalysis. Adv. Sci. 2020, 7, 2001939. [Google Scholar] [CrossRef] [PubMed]
  599. Kavil, J.; Anjana, P.M.; Joshy, D.; Babu, A.; Raj, G.; Periyat, P.; Rakhi, R.B. g-C3N4/CuO and g-C3N4/Co3O4 Nanohybrid Structures as Efficient Electrode Materials in Symmetric Supercapacitors. RSC Adv. 2019, 9, 38430–38437. [Google Scholar] [CrossRef] [Green Version]
  600. Zhou, Y.; Sun, L.; Wu, D.; Li, X.; Li, J.; Huo, P.; Wang, H.; Yan, Y. Preparation of 3D Porous g-C3N4@ V2O5 Composite Electrode via Simple Calcination and Chemical Precipitation for Supercapacitors. J. Alloys Compd. 2020, 817, 152707. [Google Scholar] [CrossRef]
  601. Klose, W.; Rincón, S. Adsorption and Reaction of NO on Activated Carbon in the Presence of Oxygen and Water Vapour. Fuel 2007, 86, 203–209. [Google Scholar] [CrossRef]
  602. Ou, M.; Zhong, Q.; Zhang, S.; Yu, L. Ultrasound Assisted Synthesis of Heterogeneous g-C3N4/BiVO4 Composites and Their Visible-Light-Induced Photocatalytic Oxidation of NO in Gas Phase. J. Alloys Compd. 2015, 626, 401–409. [Google Scholar] [CrossRef]
  603. Hu, S.; Chen, X.; Li, Q.; Li, F.; Fan, Z.; Wang, H.; Wang, Y.; Zheng, B.; Wu, G. Fe3+ Doping Promoted N2 Photofixation Ability of Honeycombed Graphitic Carbon Nitride: The Experimental and Density Functional Theory Simulation Analysis. Appl. Catal. B Environ. 2017, 201, 58–69. [Google Scholar] [CrossRef]
  604. Kong, Y.; Lv, C.; Zhang, C.; Chen, G. Cyano Group Modified g-C3N4: Molten Salt Method Achievement and Promoted Photocatalytic Nitrogen Fixation Activity. Appl. Surf. Sci. 2020, 515, 146009. [Google Scholar] [CrossRef]
  605. Hu, X.; Zhang, W.; Yong, Y.; Xu, Y.; Wang, X.; Yao, X. One-Step Synthesis of Iodine-Doped g-C3N4 with Enhanced Photocatalytic Nitrogen Fixation Performance. Appl. Surf. Sci. 2020, 510, 145413. [Google Scholar] [CrossRef]
  606. Zhang, K.; Deng, L.; Huang, M.; Tu, H.; Kong, Z.; Liang, Z.; Shao, Y.; Hao, X.; Wu, Y. Energy Band Matching WO3/B-Doped g-C3N4 Z-Scheme Photocatalyst to Fix Nitrogen Effectively. Colloids Surfaces A Physicochem. Eng. Asp. 2022, 633, 127830. [Google Scholar] [CrossRef]
  607. Liu, X.; Han, X.; Liang, Z.; Xue, Y.; Zhou, Y.; Zhang, X.; Cui, H.; Tian, J. Phosphorous-Doped 1T-MoS2 Decorated Nitrogen-Doped g-C3N4 Nanosheets for Enhanced Photocatalytic Nitrogen Fixation. J. Colloid Interface Sci. 2022, 605, 320–329. [Google Scholar] [CrossRef] [PubMed]
  608. Tao, R.; Li, X.; Li, X.; Shao, C.; Liu, Y. TiO2/SrTiO3/g-C3N4 Ternary Heterojunction Nanofibers: Gradient Energy Band, Cascade Charge Transfer, Enhanced Photocatalytic Hydrogen Evolution, and Nitrogen Fixation. Nanoscale 2020, 12, 8320–8329. [Google Scholar] [CrossRef]
  609. Camussi, I.; Mannucci, B.; Speltini, A.; Profumo, A.; Milanese, C.; Malavasi, L.; Quadrelli, P. g-C3N4-Singlet Oxygen Made Easy for Organic Synthesis: Scope and Limitations. ACS Sustain. Chem. Eng. 2019, 7, 8176–8182. [Google Scholar] [CrossRef]
  610. Ahmad, R.; Tripathy, N.; Khosla, A.; Khan, M.; Mishra, P.; Ansari, W.A.; Syed, M.A.; Hahn, Y.-B. Recent Advances in Nanostructured Graphitic Carbon Nitride as a Sensing Material for Heavy Metal Ions. J. Electrochem. Soc. 2019, 167, 37519. [Google Scholar] [CrossRef]
  611. Chen, K.; Zhang, X.-M.; Yang, X.-F.; Jiao, M.-G.; Zhou, Z.; Zhang, M.-H.; Wang, D.-H.; Bu, X.-H. Electronic Structure of Heterojunction MoO2/g-C3N4 Catalyst for Oxidative Desulfurization. Appl. Catal. B Environ. 2018, 238, 263–273. [Google Scholar] [CrossRef]
  612. Lei, G.; Zhao, W.; Shen, L.; Liang, S.; Au, C.; Jiang, L. Isolated Iron Sites Embedded in Graphitic Carbon Nitride (g-C3N4) for Efficient Oxidative Desulfurization. Appl. Catal. B Environ. 2020, 267, 118663. [Google Scholar] [CrossRef]
  613. Ma, R.; Guo, J.; Wang, D.; He, M.; Xun, S.; Gu, J.; Zhu, W.; Li, H. Preparation of Highly Dispersed WO3/Few Layer g-C3N4 and Its Enhancement of Catalytic Oxidative Desulfurization Activity. Colloids Surfaces A Physicochem. Eng. Asp. 2019, 572, 250–258. [Google Scholar] [CrossRef]
Figure 1. Number of annual publications (a) using “g-C3N4*” as a keyword since 2012, (b) using “g-C3N4*” with metal oxides (“TiO2”, “ZnO”, “WO3”, “Iron Oxide”, “Tin Oxide”, and other metal oxides) as keywords since 2012. Adapted from Scopus database, dated 1 October 2021.
Figure 1. Number of annual publications (a) using “g-C3N4*” as a keyword since 2012, (b) using “g-C3N4*” with metal oxides (“TiO2”, “ZnO”, “WO3”, “Iron Oxide”, “Tin Oxide”, and other metal oxides) as keywords since 2012. Adapted from Scopus database, dated 1 October 2021.
Nanomaterials 12 00294 g001
Figure 2. (a) Triazine and (b) tri-s-triazine (heptazine) structures of g-C3N4 (gray, blue, and white balls are carbon, nitrogen, and hydrogen, respectively).
Figure 2. (a) Triazine and (b) tri-s-triazine (heptazine) structures of g-C3N4 (gray, blue, and white balls are carbon, nitrogen, and hydrogen, respectively).
Nanomaterials 12 00294 g002
Figure 3. (a) Melamine and (b) g-C3N4 thermal analyses, Copyright © 2022 American Chemical society [31]; (c) thermogravimetric curve of annealing of melamine; (d) C/N values of melamine products at different temperatures, Copyright © 2022 Elsevier [32].
Figure 3. (a) Melamine and (b) g-C3N4 thermal analyses, Copyright © 2022 American Chemical society [31]; (c) thermogravimetric curve of annealing of melamine; (d) C/N values of melamine products at different temperatures, Copyright © 2022 Elsevier [32].
Nanomaterials 12 00294 g003
Figure 4. (a) XRD, (b) FTIR, (c) bandgaps of g-C3N4 prepared at different calcination temperatures, and (d) the photocatalytic recyclability degradation of dyes using prepared g-C3N4 at 550 °C [36].
Figure 4. (a) XRD, (b) FTIR, (c) bandgaps of g-C3N4 prepared at different calcination temperatures, and (d) the photocatalytic recyclability degradation of dyes using prepared g-C3N4 at 550 °C [36].
Nanomaterials 12 00294 g004
Figure 5. (a) The PL spectra under 325 nm excitation (b) UV-Vis spectra of g-C3N4, prepared at different temperatures, Copyright © 2022 American Chemical Society [41]; (c) The g-C3N4 in deionized water under 365 nm light; (d) the PL spectra of the g-C3N4, g-C3N4 in water and in the DI water [44].
Figure 5. (a) The PL spectra under 325 nm excitation (b) UV-Vis spectra of g-C3N4, prepared at different temperatures, Copyright © 2022 American Chemical Society [41]; (c) The g-C3N4 in deionized water under 365 nm light; (d) the PL spectra of the g-C3N4, g-C3N4 in water and in the DI water [44].
Nanomaterials 12 00294 g005
Figure 6. (a) Synthesis precursors and calcination temperature for g-C3N4 preparation; (b) synthesis procedure using cyanamide (dicyanamide), urea, thiourea, and melamine for g-C3N4 synthesis (gray, blue, red, yellow, and white balls are carbon, nitrogen, oxygen, sulfur, and hydrogen atoms, respectively).
Figure 6. (a) Synthesis precursors and calcination temperature for g-C3N4 preparation; (b) synthesis procedure using cyanamide (dicyanamide), urea, thiourea, and melamine for g-C3N4 synthesis (gray, blue, red, yellow, and white balls are carbon, nitrogen, oxygen, sulfur, and hydrogen atoms, respectively).
Nanomaterials 12 00294 g006
Figure 7. Polymerization of (a) urea and (b) thiourea into a g-C3N4 at high temperatures (gray, blue, red, yellow, and white balls are carbon, nitrogen, oxygen, sulfur, and hydrogen atoms, respectively).
Figure 7. Polymerization of (a) urea and (b) thiourea into a g-C3N4 at high temperatures (gray, blue, red, yellow, and white balls are carbon, nitrogen, oxygen, sulfur, and hydrogen atoms, respectively).
Nanomaterials 12 00294 g007
Figure 8. The two most common heterojunction types in g-C3N4–metal oxide photocatalysts: (a) type II heterojunction, (b) Z-scheme heterojunction (white and dark balls are holes and electrons, respectively).
Figure 8. The two most common heterojunction types in g-C3N4–metal oxide photocatalysts: (a) type II heterojunction, (b) Z-scheme heterojunction (white and dark balls are holes and electrons, respectively).
Nanomaterials 12 00294 g008
Figure 9. Schematic illustration of TiO2/g-C3N4 heterojunctions synthesis approach.
Figure 9. Schematic illustration of TiO2/g-C3N4 heterojunctions synthesis approach.
Nanomaterials 12 00294 g009
Figure 10. (a) Proposed mechanism for charge transfer of type II of the g-C3N4-TiO2 heterojunction interface under visible light irradiation; (b) XRD patterns; (c) FT-IR spectra of the g-C3N4-TiO2 heterojunction; (d) the pore size distribution curves; (e) UV-vis spectra of the prepared TiO2, g-C3N4, and TiO2- g-C3N4 composites, Copyright © 2022 Elsevier [120]; (f) PL emission spectrum of TiO2, g-C3N4 -TiO2, and g-C3N4 products, Copyright 2019 © American Chemical Society [122].
Figure 10. (a) Proposed mechanism for charge transfer of type II of the g-C3N4-TiO2 heterojunction interface under visible light irradiation; (b) XRD patterns; (c) FT-IR spectra of the g-C3N4-TiO2 heterojunction; (d) the pore size distribution curves; (e) UV-vis spectra of the prepared TiO2, g-C3N4, and TiO2- g-C3N4 composites, Copyright © 2022 Elsevier [120]; (f) PL emission spectrum of TiO2, g-C3N4 -TiO2, and g-C3N4 products, Copyright 2019 © American Chemical Society [122].
Nanomaterials 12 00294 g010
Figure 11. (a) TEM images of 5%-g-C3N4/TiO2 nanoparticles (inset SAED pattern); (b,c) HRTEM images of 5%-g-C3N4; (d) elemental mapping image of Ti, O, C, N elements of the 10%-g-C3N4/TiO2 composite, Copyright © 2022 Elsevier [144].
Figure 11. (a) TEM images of 5%-g-C3N4/TiO2 nanoparticles (inset SAED pattern); (b,c) HRTEM images of 5%-g-C3N4; (d) elemental mapping image of Ti, O, C, N elements of the 10%-g-C3N4/TiO2 composite, Copyright © 2022 Elsevier [144].
Nanomaterials 12 00294 g011
Figure 12. Schematic of the g-C3N4−ZnO heterojunction synthesis approach, reproduced from Reference [158].
Figure 12. Schematic of the g-C3N4−ZnO heterojunction synthesis approach, reproduced from Reference [158].
Nanomaterials 12 00294 g012
Figure 13. (a) Proposed mechanism for charge transfer of type II of the g-C3N4-ZnO heterojunction interface under visible light irradiation; (b) XRD patterns of ZnO, g-C3N4, and g-C3N4@ZnO samples; (c) UV-vis absorption spectra of ZnO, g-C3N4, and g-C3N4@ZnO composites, Copyright © 2022 Elsevier [164]; (d) Malachite Green (MG) photocatalytic activity of the g-C3N4/ZnO photocatalyst, Copyright © 2022 Elsevier [165]; (e) the pure ZnO NRAs, g-C3N4 thin film, and g-C3N4/ZnO heterostructure photocurrent response to a light on-off under visible light illumination (≥420 nm), Copyright © 2022 Elsevier [159]; (f) Nyquist plots of the ZnO, g-C3N4, and g-C3N4/ZnO catalysts. Reproduced from Reference [160].
Figure 13. (a) Proposed mechanism for charge transfer of type II of the g-C3N4-ZnO heterojunction interface under visible light irradiation; (b) XRD patterns of ZnO, g-C3N4, and g-C3N4@ZnO samples; (c) UV-vis absorption spectra of ZnO, g-C3N4, and g-C3N4@ZnO composites, Copyright © 2022 Elsevier [164]; (d) Malachite Green (MG) photocatalytic activity of the g-C3N4/ZnO photocatalyst, Copyright © 2022 Elsevier [165]; (e) the pure ZnO NRAs, g-C3N4 thin film, and g-C3N4/ZnO heterostructure photocurrent response to a light on-off under visible light illumination (≥420 nm), Copyright © 2022 Elsevier [159]; (f) Nyquist plots of the ZnO, g-C3N4, and g-C3N4/ZnO catalysts. Reproduced from Reference [160].
Nanomaterials 12 00294 g013
Figure 14. (a) Proposed structure of g-C3N4-FeOx heterojunction interface; (b) XRD patterns and (c) FT-IR spectra of g-C3N4, α-Fe2O3, and α-Fe2O3/g-C3N4 composites, Copyright © 2022 Elsevier [220]; (d) UV-Vis diffuse reflection absorption spectra; (e) PL spectra of CN and xFe-CN samples with an excitation wavelength of 380 nm, Copyright © 2022 Elsevier [221]; (f) single-wavelength photocurrent response of F, 15CN-F and 15CN-6Al-F, Copyright © 2022 Elsevier [222].
Figure 14. (a) Proposed structure of g-C3N4-FeOx heterojunction interface; (b) XRD patterns and (c) FT-IR spectra of g-C3N4, α-Fe2O3, and α-Fe2O3/g-C3N4 composites, Copyright © 2022 Elsevier [220]; (d) UV-Vis diffuse reflection absorption spectra; (e) PL spectra of CN and xFe-CN samples with an excitation wavelength of 380 nm, Copyright © 2022 Elsevier [221]; (f) single-wavelength photocurrent response of F, 15CN-F and 15CN-6Al-F, Copyright © 2022 Elsevier [222].
Nanomaterials 12 00294 g014
Figure 15. (a) Proposed structure of g-C3N4-WO3 heterojunction interface; (b) XRD patterns and (c) FT-IR spectra of obtained samples (x in WO3/g-C3N4 refer to the mass ratio of WO3 to g-C3N4), Copyright © 2022 Elsevier [251]; (d) UV-DRS spectra of pure g-C3N4, WO3 and WO3/g-C3N4 heterojunctionp; (e) PL emission spectra of g-C3N4 and 18.6 wt % WO3/g-C3N4 composite, Copyright © 2022 Elseviers [254].
Figure 15. (a) Proposed structure of g-C3N4-WO3 heterojunction interface; (b) XRD patterns and (c) FT-IR spectra of obtained samples (x in WO3/g-C3N4 refer to the mass ratio of WO3 to g-C3N4), Copyright © 2022 Elsevier [251]; (d) UV-DRS spectra of pure g-C3N4, WO3 and WO3/g-C3N4 heterojunctionp; (e) PL emission spectra of g-C3N4 and 18.6 wt % WO3/g-C3N4 composite, Copyright © 2022 Elseviers [254].
Nanomaterials 12 00294 g015
Figure 16. (a) Proposed structure of g-C3N4-SnO2 heterojunction interface; (b) XRD patterns and (c) FT-IR spectra of SnO2, graphene-like C3N4, SnO2/g-C3N4 composites (reproduce from Reference [295]); (d) UV-vis absorption spectra of g-C3N4-based samples (muscovite sheet(x)/SnO2/g-C3N4 (MSx/CN), which x refers to the mass of MS powder), Copyright © 2022 Elsevier [298]; (e) PL spectra of g-C3N4 and SnO2 quantum dots-g-C3N4 nanocomposite, Copyright © 2022 Elsevier [299]; (f) EIS Nyquist plots of the bare SnO2, g-C3N4, and SnO2-g-C3N4 structure, Copyright © 2022 Elsevier [300].
Figure 16. (a) Proposed structure of g-C3N4-SnO2 heterojunction interface; (b) XRD patterns and (c) FT-IR spectra of SnO2, graphene-like C3N4, SnO2/g-C3N4 composites (reproduce from Reference [295]); (d) UV-vis absorption spectra of g-C3N4-based samples (muscovite sheet(x)/SnO2/g-C3N4 (MSx/CN), which x refers to the mass of MS powder), Copyright © 2022 Elsevier [298]; (e) PL spectra of g-C3N4 and SnO2 quantum dots-g-C3N4 nanocomposite, Copyright © 2022 Elsevier [299]; (f) EIS Nyquist plots of the bare SnO2, g-C3N4, and SnO2-g-C3N4 structure, Copyright © 2022 Elsevier [300].
Nanomaterials 12 00294 g016
Figure 17. (a) Hydrogen production rate of different samples as a function of the amount of methanol added as a sacrificial agent under solar light irradiation, Reproduced from Reference [403]; (b) schematic image of the ZnO/C3N4 QDs synthesis on FTO using electrodeposition and a dip-coating method, Copyright © 2022 Elsevier [406].
Figure 17. (a) Hydrogen production rate of different samples as a function of the amount of methanol added as a sacrificial agent under solar light irradiation, Reproduced from Reference [403]; (b) schematic image of the ZnO/C3N4 QDs synthesis on FTO using electrodeposition and a dip-coating method, Copyright © 2022 Elsevier [406].
Nanomaterials 12 00294 g017
Figure 18. (a) Hydrocarbon generation of hollow g-C3N4 (H-g-C3N4), hollow CeO2 (H-CeO2), and g-C3N4@CeO2 for 4 h illumination; (b) photoluminescence emission spectra of H-g-C3N4, H-CeO2 references, and g-C3N4@CeO2 heterojunction with different ratios. Copyright © 2022 Elsevier [435].
Figure 18. (a) Hydrocarbon generation of hollow g-C3N4 (H-g-C3N4), hollow CeO2 (H-CeO2), and g-C3N4@CeO2 for 4 h illumination; (b) photoluminescence emission spectra of H-g-C3N4, H-CeO2 references, and g-C3N4@CeO2 heterojunction with different ratios. Copyright © 2022 Elsevier [435].
Nanomaterials 12 00294 g018
Figure 19. (a) TG and (b) DTG curves of pure ammonium perchlorate (AP), AP coupled with g-C3N4, CeO2, and g-C3N4/CeO2 heterojunction, Copyright © 2022 Elsevier [460].
Figure 19. (a) TG and (b) DTG curves of pure ammonium perchlorate (AP), AP coupled with g-C3N4, CeO2, and g-C3N4/CeO2 heterojunction, Copyright © 2022 Elsevier [460].
Nanomaterials 12 00294 g019
Figure 20. Different kinds of g-C3N4–metal oxide-based sensors for materials detection.
Figure 20. Different kinds of g-C3N4–metal oxide-based sensors for materials detection.
Nanomaterials 12 00294 g020
Figure 21. (A) Photocurrent responses of (a) FTO, (b) FTO/TiO2, (c) FTO/TiO2/g-C3N4, (d) FTO/TiO2/g-C3N4/CdS, (e) FTO/TiO2/g-C3N4/CdS/capture-DNA3/MCH, (f), FTO/TiO2/g-C3N4/CdS/capture-DNA3/MCH/DNA2-CdSe, (g) FTO/TiO2/g-C3N4/CdS/capture-DNA3/MCH/CdSe, Copyright © 2022 American Chemical Society [557]; (B) Nyquist plots of ZnO, g-C3N4/ZnO composites and g-C3N4 nanosheets electrodes, Copyright © 2022 American Chemical Society [559].
Figure 21. (A) Photocurrent responses of (a) FTO, (b) FTO/TiO2, (c) FTO/TiO2/g-C3N4, (d) FTO/TiO2/g-C3N4/CdS, (e) FTO/TiO2/g-C3N4/CdS/capture-DNA3/MCH, (f), FTO/TiO2/g-C3N4/CdS/capture-DNA3/MCH/DNA2-CdSe, (g) FTO/TiO2/g-C3N4/CdS/capture-DNA3/MCH/CdSe, Copyright © 2022 American Chemical Society [557]; (B) Nyquist plots of ZnO, g-C3N4/ZnO composites and g-C3N4 nanosheets electrodes, Copyright © 2022 American Chemical Society [559].
Nanomaterials 12 00294 g021
Figure 22. (a) Plots of bacterial counts of g-C3N4/ZnO/cellulose-0.45 (0.45 corresponds to the mass of g-C3N4 addition), ZnO/cellulose, ZnO, g-C3N4, and blank sample against S. aureus and (b) E. coli; (c) bacterial counts of g-C3N4/ZnO/cellulose composites against S. aureus and (d) E. coli. Reproduced from Reference [202].
Figure 22. (a) Plots of bacterial counts of g-C3N4/ZnO/cellulose-0.45 (0.45 corresponds to the mass of g-C3N4 addition), ZnO/cellulose, ZnO, g-C3N4, and blank sample against S. aureus and (b) E. coli; (c) bacterial counts of g-C3N4/ZnO/cellulose composites against S. aureus and (d) E. coli. Reproduced from Reference [202].
Nanomaterials 12 00294 g022
Table 1. The electrical properties and application of doped g-C3N4.
Table 1. The electrical properties and application of doped g-C3N4.
Doping ElementEcEgEvApplicationFurther ExplanationRef
Phosphorus−1.11 eV2.55 eV1.44 eVCatalytic aromatic alcoholsThe presence of P enhances the aldehyde selectivity.[73]
−0.33 eV2.58 eV2.25 eVPhotocatalytic hydrogen evolution The g-C3N4 tube doped with P improves light absorption. The quantum efficiency of the P-doped g-C3N4 tube is 4.7 and 22.4 times higher than that of the g-C3N4 tube and bulk g-C3N4.[74]
−1.34 eV2.79 eV1.44 eVPhotocatalytic CO2 conversionThe P-modified g-C3N4 demonstrates the highest photocatalytic efficiency.[75]
−1.17 eV2.69 eV1.52 eVPhotocatalytic hydrogen evolutionThe P-doped structure has a high efficiency in accordance with the recombination, migration, and separation of electron-hole pairs.[76]
Sulfur−1.04 eV2.92 eV1.88 eVPhotocatalytic nitrogen fixationSulfur enhances the adsorption and activation of N2 molecules of g-C3N4 porous nanosheets and uses for photocatalytic nitrogen fixation.[77]
−1.23 eV2.80 eV1.57 eVPhotocatalytic hydrogen evolutionThe H2 generation rate of N-doped MoS2 and S-doped g-C3N4 is about 23 and 38 times higher than that of pure SCN and NMS with 28.8 μmol/g/h and 17.4 μmol/g/h, respectively.[78]
−1.3 eV2.67 eV1.34 eVPhotocatalytic hydrogen evolutionThe BiPO4/S-C3N4 improves photocatalytic activity by facilitating carrier transportation.[79]
−1.3 eV2.69 eV1.39 eVPhotocatalytic bisphenol degradation Ag–S-C3N4 enhanced the photocatalytic activity since Ag has a great electron storage ability.[80]
−1.32 eV2.66 eV1.34 eVPhotocatalytic hydrogen evolutionSulfur promotes the photocatalytic ability of hydrogen evolution about four times higher than the bulk g-C3N4.[81]
Oxygen−0.88 eV2.61 eV1.73 eVPhotocatalytic CO2 reductionThe porous O-doped graphitic carbon nitride reveals enhanced photocatalytic activity.[82]
−0.76 eV2.57 eV1.84 eVPhotocatalytic hydrogen evolutionThis result of the band edge value is related to the 1.1% oxygen content mass percentage.[83]
−0.37 eV2.53 eV2.15 eVThis result of the band edge value is related to the 2.3% oxygen content mass percentage.
−1.08 eV2.93 eV1.85 eVPhotocatalytic hydrogen evolution and 2,4-dinitrophenolThe oxygen dopant with Pt exhibits excellent photocatalytic hydrogen evolution in overall water splitting with 29.6 μmol/(g·h), and O-g-C3N4 NR reached up to approximately 100% removal efficiency of 2,4-dinitrophenol within 75 min.[84]
1.51 eV2.70 eV−1.19 eVphotocatalytic water splittingThis band edge value is related to CN-x = 0 (x refers to the quantity of citric acid (gr)).[85]
1.50 eV2.62 eV−1.16 eVThis band edge value is related to CN-0.2.
1.46 eV2.52 eV−1.06 eVThis band edge value is related to CN-0.4.
1.46 eV2.49 eV−1.03 eVThis band edge value is related to CN-0.6.
Carbon−1.13 eV2.54 eV1.41 eVThermal oxidation etching processC-doped g-C3N4 improves the catalytic activity by extending the visible light absorption.[86]
Boron−0.8 eV2.8 eV2 eVPhotocatalytic Oxygen evolution, Cr(VI) reductionThis structure is used for Cr(VI) reduction and O2 generation simultaneously.[87]
Nitrogen−0.5 eV2.4 eV1.9 eVPhotocatalytic tetracycline degradationThis band edge value is related to the nitrogen-doped g-C3N4 used for Photocatalytic tetracycline degradation.[88]
−0.6 eV2.5 eV1.9This band edge value is related to the nitrogen-doped g-C3N4 nanosheets.
−0.33 eV1.82 eV1.49 eVPhotocatalytic phenol degradation N-doped g-C3N4 possesses a narrow bandgap since the N atom introduced an inter-bandgap, resulting in the redshift in the absorption UV-vis peak.[89]
MetalNa−1.16 eV2.77 eV1.36 eV17α-ethynylestradiol mineralizationThe photostable Na doped g-C3N4 content causes the photoabsorption enhancement. [70]
K−1.08 eV2.72 eV1.64 eVPhotocatalytic CO2 reductionK content causes defects leading to improving catalytic activity by reducing the electron-hole recombination.[90]
Ti−1.02 eV2.5 eV1.48 eVPhotocatalytic enhancementTi-doped g-C3N4 caused narrower bandgap and reduced carrier recombination resulting in higher absorption.[91]
Mn−0.59 eV2.56 eV1.97 eVPhotocatalytic methylene blue degradation The Mn-doped g-C3N4 nanoribbon reveals a great potential photocatalytic agent for water splitting coupling with MB degradation.[92]
Ag-2.60 eV-Photocatalytic oxidation of methylene blueThe higher Ag content leads to the lower bandgap of the structure, and a lower recombination rate is observed in the Ag-doped g-C3N4.[93]
Fe−1.10 eV2.50 eV1.40 eVEnvironmental pollution controlFe3+ with nitrogen in heptazine forms a σ-π bond and can accelerate the electron-hole separation.[94]
Co−0.36 eV2.62 eV2.26 eVPhoto-electrochemical water oxidationCo-doped g-C3N4 reduces the electron-hole recombination rate and demonstrates promising photocurrent and electrical conductivity.[95]
Co-dopedP, O−0.80 eV2.30 eV1.50 eVPhotocatalytic fluoroquinolone antibiotics degradationThe degradation rate of enrofloxacin was 6.2 times higher for phosphorus and oxygen co-doped graphitic carbon nitride (POCN) than g-C3N4.[96]
P, S-2.6 eV-Photocatalytic hydrogen evolutionThe high photocatalytic activity can be observed because of the synergic impact of P and S co-doping.[97]
B, F-2.72 eV-Photocatalytic hydrogen evolutionB, F co-doped g-C3N4 improves the charge generation and the separation efficiency.[98]
Na, O-2.72 eV-Photocatalytic hydrogen evolutionNa, O co-doped g-C3N4 reveals that photocatalytic H2 production activity was seven-fold improved by enhancing absorption of UV-vis spectra. [99]
Table 2. A list of selected works on g-C3N4–metal oxide-based photocatalytic water splitting.
Table 2. A list of selected works on g-C3N4–metal oxide-based photocatalytic water splitting.
PhotocatalystType of HeterojunctionSource of LightHighest Photocatalytic RateRef
TiO2-g-C3N4Type IIAsahi Spectra Hal-320 (300 mW cm−2) with a 420 nm cut off filter (λ > 420 nm)3.6 μmol·h−1[115]
TiO2-g-C3N4Type IIXenon lamp with a 320 nm cut off filter (λ > 320 nm)76.25 μmol·h−1[414]
C-doped TiO2-g-C3N4Type II300 W Xe lamp (PLS-SXE300)
with a 420 nm cutoff filter (λ > 420 nm)
35.6 μmol·g−1·h−1[128]
TiO2-g-C3N4 decorated by Co-PiType II300 W Xe lamp coupled with a monochromator-
TiO2 nanodots/g-C3N4S-scheme300 W Xe lamp (LANPU)
with a 300 nm cutoff filter (λ > 300 nm)
The H2 and O2 evolution rate is 1318.3 and 638.7 μmol g−1, respectively, (roughly as same as the stoichiometric ratio of evolved H2 to O2)[415]
ZnO-g-C3N4Type IIPLS-SXE-300C lamp with an UV light intensity of 34 mW/cm2 and visible-light intensity of 158 mW/cm2-[416]
N-doped ZnO-g-C3N4Z-schemePLS-SXE-300C UV lamp with a 420 nm cut off filter (λ > 420 nm)152.7 μmol·h−1[417]
g-C3N4-WO3-300-W Xe lamp (PLS-SXE300) with a 420 nm cutoff filter (λ > 420 nm)963 μmol·g−1·h−1[141]
O-g-C3N4/WO3Z-scheme300 W Xenon lamp with a 420 nm cutoff filter (λ > 420 nm)15,142 μmol·g−1[265]
S-Cu2O/g-C3N4Z-scheme300 W Xe lamp with a 420 nm cutoff filter (λ > 420 nm)24.83 μmol·h−1[418]
BiO2/g-C3N4Type II
Z-scheme
500 W Xe lamp with a 420 nm cutoff filter (λ > 420 nm)8,542 μmol·g−1[419]
g-C3N4/BiYO3Type II-37.6 μmol·g−1·h−1[420]
g-C3N4/LaxCo3-xO4-300 W Xe lamp with a 420 nm cutoff filter (λ > 420 nm)63.12 μmol·h−1[421]
Fe2O3/g-C3N4Z-scheme350 W Xe lamp with a 420 nm cutoff filter (λ > 420 nm)398.0 μmol·g−1·h−1[210]
Mn3O4/g-C3N4p-n heterostructure300 W Xenon lamp with a 420 nm cutoff filter (λ > 420 nm) (PLS-SXE300D/300DUV, Beijing Perfectlight)The H2 and O2 evolution rate is 3300 and 654 μmol g−1·h−1, respectively.[422]
g-C3N4/Nitrogen-Doped Carbon Dots/WO3-300 W Xenon lamp with a 420 nm cutoff filter (λ > 420 nm) (CEL-HXF 300)3.27 mmol g–1 h–1[423]
Mn3O4/g-C3N4-300 W Xenon lamp source (PLS-SXE300D/300DUV)2700 mmol·g−1·h−1[422]
NiO/g-C3N4Type IIXe lamp with a 420 nm cutoff filter (λ > 420 nm)1.41 mmol·h−1[424]
In2O3/g-C3N4Type II300 W Xe lamp with a 420 nm cutoff filter (λ > 420 nm)0.99 mmol·h−1[425]
MoO3-x-g-C3N4Z-scheme300 W Xe lamp with a 420 nm cutoff filter (λ > 420 nm)22.8 mmol·h−1[426]
ZnO/Au/g-C3N4Z-scheme150 W Xenon arc lamp with a 420 nm cutoff filter (λ > 420 nm)3.69 μmol h−1 cm−2[397]
d-Ti3C2/TiO2/g-C3N4-300 W Xe lamp with a 420 nm cutoff filter (λ > 420 nm)1.62 mmol·h−1g−1[396]
TiO2/g-C3N4Type II450 W high-pressure mercury lamp22.4 mol·h−1[402]
TiO2-WO3-g-C3N4--286.6 mmol·h−1[427]
TiO2/Ti3C2/g-C3N4-300 W Xe lamp2592 mmol·g−1[428]
Table 3. A list of CO2 reduction applications of the g-C3N4–metal oxide-based photocatalytic.
Table 3. A list of CO2 reduction applications of the g-C3N4–metal oxide-based photocatalytic.
PhotocatalystType of HeterojunctionSource of LightHighest Photocatalytic RateRef
NiO-g-C3N4Type II300 W Xenon-arc lamp4.17 μmol·g−1·h−1[436]
g-C3N4 foam-Cu2OZ-scheme350–780 nm lamp8.182 μmol·g−1·h−1
(CO revolution)
[437]
NiMoO4-g-C3N4Z-scheme-7238 μmol·g−1·h−1[438]
CeO2-g-C3N4Type II300 W of Xenon-arc lamp0.590 μmol·h−1
(CO evolution)
[439]
ZnO/Au/g-C3N4Z-scheme300 W UV-Vis lamp689.7 μmol/m2
(CO evolution)
[440]
ZnO/g-C3N4Z-scheme300 W xenon light source with a 420 nm cutoff filter (λ > 420 nm)~72.24 μmol·g−1[435]
TiO2/g-C3N4Type II8 W UV lampThe highest CH4 and CO yields of 72.2 and 56.2 μmol g−1[431]
Nb doped TiO2/g-C3N4Z-scheme1000 W Xe lampThe CH4, CO, O2, HCOOH generation rate in the presence of 50Nb-TiO2/50 g-C3N4 is 562, 420, 1702, 698 μmol h−1 g−1, respectively.[432]
ZnO/g-C3N4Z-scheme350 W Xe lampThe CH3OH production rate was 1.32 μmol h−1 g−1[429]
ZnO/g-C3N4Type II300 W xenon lamp with a 420 nm cutoff filter (λ > 420 nm)H2, CH4, and CO production rates of 22.7 μmol·g-Cat−1·h−1, 30.5 μmol·g-Cat−1·h−1, and 16.8 μmol·g-Cat−1·h−1[433]
ZnO/g-C3N4Type II350 W Xe arc lamp45.6 mol·g-Cat−1·h−1[441]
TiO2/g-C3N4Type II450 W Xe lamp22.5 μmol·g−1 and 70 μmol·g−1 for CO and CH4 yield, respectively[442]
g-C3N4/3D ordered microporous (3DOM)-WO3Z-schemevisible light (λ ≥ 420 nm)48.7 μmol g−1 h−1[443]
NiTO3/g-C3N4Z-scheme300 W xenon lamp with a 420 nm cutoff filter (λ > 420 nm)The highest yield of CH3OH production is 13.74 μmol∙g−1∙h−1[444]
Table 4. the photodegradation application of g-C3N4–metal oxide-based heterojunctions.
Table 4. the photodegradation application of g-C3N4–metal oxide-based heterojunctions.
PhotocatalystType of HeterojunctionSource of LightApplicationHighest Photocatalytic RateStabilityRef
TiO2-g-C3N4Z-scheme300 W Xenon arc lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of Rhodamine B and tetracycline hydrochlorideThe RhB removal rates for 5 layer TiO2, 3, 5, 7 layers g-C3N4 (0.5)/TiO2 were 5.1%, 17.9%, 31.2%, and 22.6%, respectively-[503]
P/O co-doped g-C3N4/anatase TiO2Z-scheme350 W Xenon-arc lamp as a light source with a 420 nm cutoff filter (λ > 420 nm)Degradation of enrofloxacin~98.5%1 h[504]
TiO2@ g-C3N4Z-scheme100-W xenon lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of RhB95.68% [505]
MoS2-g-C3N4@TiO2-350 W Xenon lampDegrdation of Methylene Blue 97.551 h[506]
g-C3N4 and polyaniline-co-modified TiO2-xenon lamp containing an optical filterDegradation of tetrabromobisphenol A92.42%16 h[507]
N-TiO2/O-doped N vacancy g-C3N4Type II
Z-scheme
Lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of tetracycline hydrochloride and Cr(VI)TC-HCl and Cr(VI) removal efficiency is 79.9% and 89.5%, respectively-[508]
TiO2@g-C3N4Type II300 W xenon lampDegradation of tetracycline antibioticTiO2@g-C3N4 photocatalyst shows the The highest tetracycline degradation rate is 2.2 mg/min, which is 2 times higher than that of TiO2 and 2.3 times higher than that of bulk g-C3N4.-[509]
TiO2/g-C3N4/persulfate (PS)Type II300 W xenon lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of micropollutant (phenol, bisphenol A and carbamazepine99.3%. [510]
TiO2 nanowire/g-C3N4 nanosheet/graphene (G) heterostructures-300 W xenon lampDegradation of nitrobenzene97%4 h[119]
Ti3+ and O doped TiO2/g-C3N4Type II30 W cold visible light-emitting diodeDegradation of Rhodamine BThe photodegradation reaction rate constant based on this heterojunction is 0.0356 min−1, which is 3.87 and 4.56 times higher than those of pristine Ti3+-TiO2 and g-C3N4, respectively.-[125]
TiO2/g-C3N4Type II-Degrdation of ciprofloxacin (CIP)68.1%3 h[511]
ZnO-g-C3N4Z scheme500 W Xe lamp, with a 420 nm cutoff filter (λ > 420 nm)Degradation of Methylene Blue (MB)75%3 h[512]
ZnO-g-C3N4Z scheme300 W xenon lampDegradation of cephalexin oxidation98.9%1 h[513]
djembe-like ZnO- g-C3N4-150 W Xenon light sourcesDegradation of MB and RhBMB and RhB degradation efficiency are ~95% and ~97%, respectively.50 min[514]
WO3-g-C3N4 300 W Xenon lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of tetracycline90.54%1 h[515]
WO3-g-C3N4Z-scheme450 W xenon lampsDegradation of AO7100%75 min[516]
WO3-g-C3N4Z-scheme35-W Xe lamp with a radiation intensity of 1380 μW∙cm2Degradation of orange G98%1 h[517]
Ag-WO3/g-C3N4Z-scheme500 W Xe lamp (Beijing Bofei Technology. Co. Ltd., Beijing, China)degradation of oxytetracycline hydrochloride97.741 h[273]
WO3@g-C3N4@MWCNTZ-scheme5 mW cm−2 Xe lamp (SANEI electronics-JAPAN)Degradation of tetracycline79.54%2 h[518]
g-C3N4/WO3Z-scheme-Degradation of nitenpyramThe photocatalyst rate constant is 0.036 min−1 which is about 1.7 and 25 times higher than that of pure g-C3N4 and WO3, respectively.-[519]
Bi2O3- g-C3N4Z-scheme350 W xenon lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of Rhodamine B98% 80 min[520]
Bi2O3- g-C3N4Z-scheme250 W Xenon lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of tetracycline80.2%50 min[521]
g-C3N4-CeO2Type II300W Xe arc lampDegradation of antibiotic doxycycline hydrochloride84%1 h[522]
Shuttle-like g-C3N4-CeO2-300 W Xe arc lampDegradation of norfloxacin88.6%1 h[523]
Kaolin/CeO2/g-C3N4-500 W Xenon lamp with a 420 nm cutoff filter (λ > 420 nm)Removal of ciprofloxacin90%150 min[524]
Fe3O4/CeO2/g-C3N4-300 W Xe lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of tetracycline hydrochloride96.63%3 h[525]
Au/g-C3N4 nanosheets/CeO2Z-scheme500 W Xe lamp with a 400 nm cutoff filter (λ > 400 nm)Reduction of hexavalent chromium
Oxidation of oxytetracycline hydrochloride
88.2%
95.1%
150 min[411]
Co3O4-g-C3N4-250 W Xe lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of Methyl Orange100%3 h[526]
NiO-g-C3N4Type II500 W Xe-lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of Methylene Blue6.3 wt. % NiO loading shows a 2.3 times higher MB degradation rate than that of the pristine g-C3N4.80 min[527]
V2O5-g-C3N4Z-scheme500 W Xenon lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of Congo Red and Cr (VI) reduction-.90 min[528]
MoO3-g-C3N4Z-scheme150 W-Xe lamp having 1.5 AM filter which allows wavelength
larger than 400 nm for the visible light-based catalytic reaction
Degradation of Rhodamine B93%3 h[529]
MoO3-g-C3N4Z-scheme500W Xenon lampDegradation of Rhodamine B100%10–15 min[530]
MoO3-g-C3N4Z-schemevisible light (λ > 420 nm)Degradtion of tetracycline85.9%100 min[531]
BiMoO6-g-C3N4Z-scheme300 W Xe lampDegradation of ciprofloxacin~10030 min[532]
g-C3N4/SnO2S-schemeA 300 W Osram, 230 V with a 420 nm cut-off filter as used as the visible light source.Degradation of NO44.17%30 min[533]
g-C3N4-NiOZ-scheme30 W LED-light sourceDegradation of Methyl Orange (MO)96.8%2 h[461]
ZnO/g-C3N4Type II150 W Xe lampDegradation of MB and RhBThe MB and RhB degradation was ~95% and ~97%, respectively.50 min[514]
g-C3N4/CuOx-350W Xe lampDegradation of Methyl Orange (MO)~62.5%.70 min[534]
WO3/g-C3N4Z-scheme300 W Xe arc lampDegradation of sulfamethoxazole91.7%4 h[453]
g-C3N4 Nanosheets/ZnO-500 W Xe lampPhotocatalytic reduction of aqueous chromium(VI)70%4 h[535]
ZnO/g-C3N4-500 W xenon lampPhotodegradation of Direct Blue 199 (DB)99%100 min[536]
ZnO/g-C3N4-A 150 W xenon lampDegradation of tetracycline hydrochloride97%30 min[452]
MoS2/Al2O3/g-C3N4-150 W tungsten halogen lamp with a 420 nm cutoff filter (λ > 420 nm)Degradation of crystal violet (CV)97.3%.90 min[537]
g-C3N4/ZnOZ-scheme300 W Xenon lamp cutoff filter with a 420 nm cutoff filter (λ > 420 nmDegradation of 4-chlorophenol~ 95%1 h[538]
Gd2O3 NPs@g-C3N4-300 W Xe lamp with a 420 nm cutoff filter (λ > 420 nm) (PLS-SXE300, Beijing Perfectlight Technology Co., Ltd., Beijing, China) Degradation of Methyl Orange (MO)
Methyl Blue (MB)
Rhodamine B
72.4%
95.5%
100%
2 h[539]
MoO3/g-C3N4/peroxydisulfate (PDS)Z-scheme350 W Xenon lamp with a 420 nm cutoff filter (λ > 420 nm)degradation of
ofloxacin (OFLX)
94.4%2 h[540]
ZnO/g-C3N4Z-scheme300 W Xe lampDegradation of cephalexin98.9%1 h[513]
Ce2O/Bi2O3/g-C3N4Z-scheme75 W halogen lampDegradation of Malachite green and Rose Bengal-1 h[541]
CuO/ZnO/g-C3N4Z-scheme400 W hallow lampDegradation of Methylene Blue and ammonia-nitrogenThe MB and ammonia-nitrogen degradation efficiency are
~98% in 45 min and 91% in 6 h.
45 min, and 6 h[542]
Fe3O4/ZnO/g-C3N4Type II23 W white LEDDegradation of pantoprazole97.09%.90 min[543]
Fe3O4/TiO2/g-C3N4-500 W Xenon Lamp with a 420 nm cutoff filter (λ > 420 nmDegradation of Rhodamine B (RhB) and Methylene Orange (MO)The photocatalyst shows the RhB, and MO degradation efficiency is ~96.4% in 80 min and 90% in120 min.8 min and 120 min[544]
g-C3N4/ZnO/TiO2-1000 W xenon lamp irradiation system equipped with a 410 nm cutoff filter under room temperature.Degradation of p-toluenesulfonicacid (p-TSA)100%60 min[545]
ZnO/Ag2O/g-C3N4-high pressure xenon short arc lamp with the light intensity of 100 mW cm−2Degradation of ciprofloxacin97.4%48 min[196]
WO3/TiO2@g-C3N4-500 W metal halide lampRemoval of acetylsalicylate (aspirin) and methyl-theobromine (caffeine)98%90 min[546]
TiO2@g-C3N4/Co3O4-300 W xenon lampDegradation of tetracycline (TC) and Methylene Orange (MO)The TC (10 mg/L) and MO (25 mg/L) degradation efficiency are 91.6% and 97.8%, respectively.1 h[547]
g-C3N4/Bi2O3/TiO2-Xe-lamp light source with a 420 nm cutoff filter (λ > 420 nmDegradation of Methylene Blue (MB)The MB removal efficiency is 77.5%.3 h[548]
SnO2/ZnO@g-C3N4-300 W xenon lampDegradation of Rhodamine B dye and H2 production99%1 h[494]
g-C3N4/NiO/ZnO/Fe3O4--removal of esomeprazole95.05 ± 1.72%70 min[513]
Table 5. The list of applications of g-C3N4–metal oxide heterojunction used for detection.
Table 5. The list of applications of g-C3N4–metal oxide heterojunction used for detection.
PhotocatalystApplicationExplanationRef
MoO3-g-C3N4Detection of FurazolidoneThe high electroactive surface area (0.3788 cm2), as well as enhanced heterogeneous electron transfer rate (K°eff = 4.91 × 10−2 cm·s−1 can detect Furazolidone with low limit of detection (LOD) (1.4 nM) with a working range of 0.01–228 μM.[528]
Co3O4-g-C3N4Detection of environmental phenolic hormonesThis composite showed a wide detection range and a low limit of detection LOD (10−9 mol L−1)[565]
g-C3N4-Fe3O4Determination of Tramadol in Human Biological FluidsLOD of this composite is ~0.1 μΜ[566]
V2O5-g-C3N4Detection of folic acidThe sensitivity, the LOD, and noise-to-signal ratio of the sensor is 19.02 μA mM−1 cm−2, and 0.00174 μM, and 3 (S/N = 3), respectively[567]
g-C3N4-NiODetection of quercetinThe dynamic range and LOD of g-C3N4-NiO for sensing quercetin is 10 nM to 250 and 0.002 μM, respectively[568]
Cu2O-g-C3N4Humidity sensorThe response time and recovery time were 180–200 s and 5–10 s, respectively.[569]
NiO-Co3O4-g-C3N4Detection of tetrabromobisphenol-AThis structure showed a LOD of ~0.1 mmol L−1[570]
ZnO @ g-C3N4Detection of CCRF-CEM cellsThe LOD of this compsite is ~20 cell/mL[571]
ZnO flower-rod/g-C3N4-gold nanoparticleDetection of carcinoembryonic antigen (CEA)The PEC aptasensor for CEA determination is from 0.01 to 2.5 ng·mL−1 with detection of 1.9 pg·mL−1.[572]
g-C3N4/ZnOEthanol sensingCompared to the ZnO, the g-C3N4/ZnO-8% composites revealed an excellent response ~60 orders of magnitude at room temperature.[559]
ZnO/g-C3N4NO2 sensingThe response, recovery time, and LOD of ZnO/g-C3N4-10 wt % are 142, 190 s, and 38 ppb, respectively.[561]
SnO2/g-C3N4Ethanol gas SensingThe composite with 7 wt % g-C3N4 content exhibited a promising gas sensing property to ethanol, which has better response and selectivity than that of the pure SnO2 based sensor.[321]
g-C3N4/ZnOCH4 SensingThe higher active sites can be obtained in this structure due to the larger specific surface area leading to the great response toward 1000 ppm CH4.[560]
g-C3N4-Mn3O4H2S sensorThe LOD (S/N = 3, LOD) is 0.13 μg mL−1[573]
α-Fe2O3/g-C3N4H2S sensorThe linear detection range and detection limit of the H2S gas sensor were 0.88–7.01 μg mL−1(r = 0.998) and 0.5 μg mL−1(S/N = 3), respectively.[562]
O vacancy WO2.9/g-C3N44-nitrophenol SensingCompared to other research works, this heterojunction showed the linear range of 0.4–100 μmol/L and a lower detection limit of 0.133 μmol/L.[574]
b-Bi2O3/g-C3N4Detection of chlorpyrifosThe linear detection range and detection limit of this sensor were 0.1–80 ng mL−1 a03 μg mL−1, respectively.[575]
Fe3O4/Bi2O3/g-C3N4Determination of Cd2+ and Pb2+The minimum quantity for Cd2+ and Pb2+ that can be detected is 3 × 10−9 and 1 × 10−9 mol/L, respectively.[576]
WO3/g-C3N4Detection of phosmetThe limit of detection (LOD) and limit of quantification (LOQ) of this device was calculated 3.6 nM and 11.2 nM, respectively.[577]
WO3/g-C3N4/MnO2Detection of oxytetracycline cathodicThis sensor exhibited a wide detection range from 1 pM to 150 nM and a low detection limit of 0.1 pM.[578]
CuO-g-C3N4Aflatoxin B1 sensingThe limit of detection of 6.8 pg mL−1 for AFB1.[579]
Table 6. Number of researches on the disinfection ability of the g-C3N4 metal oxide-based nanomaterials.
Table 6. Number of researches on the disinfection ability of the g-C3N4 metal oxide-based nanomaterials.
PhotocatalystSource of LightApplicationHighest Photocatalytic RateRef
g-C3N4/TiO2/AgXe lamp with a 420 nm cutoff filter (λ > 420 nm)Bactericidal efficiency against E. colithe optimal bacterial inhibition of g-C3N4/TiO2/Ag was 84%[102]
g-C3N4/Cu2O300 W xenon lamp with 400 nm cutoff filter (λ > 400 nm)Inactivation efficiencies of E. coli as well as Fusarium graminearumg-C3N4 with 45%Cu2O composition revealed the inactivation efficiencies of ~7 log E. coli[584]
TiO2/g-C3N4/SnO2300 W xenon lamp with 420 nm cutoff filter (λ > 420 nm)Bactericidal efficiency against E. coliTiO2/g-C3N4/SnO2 structure showed a good E. coli disinfection efficiency under visible and UV irradiation, with ~−6.7 log E. coli and −8.2 log E. coli, respectively.[314]
α-Fe2O3/CeO2 decorated g-C3N4500 W Xe light with the cut-off filter of ~420 nm (λ > 420 nm)Antibacterial activities against S. aureus (G+) and E. coli (G−) bacteriaα-Fe2O3/CeO2 decorated g-C3N4 exhibited perfect antibacterial to E. coli and S.aureus activity with the maximum zone of inhibition (ZOI) of 11 ± 0.5, 12 ± 0.[585]
g-C3N4-m-Bi2O4500 W halogen lamp with UV-cut off filter (λ > 420 nm)Bactericidal efficiency E. coli and S.aureus bacteriaThe ZOI value for E.coli and S. aureus bacterial strains was ~11 ± 0.5 mm and 12-13±0.5.[586]
Ag/ZnO/g-C3N4300 W xenon arc lampBactericidal efficiency against E. coliThe E. coli disinfection efficiency of Ag/ZnO/g-C3N4 structure is ~7.4 log E. coli.[190]
Ag/AgO-g-C3N4100 W tungsten lampBactericidal efficiency against E. coliThe 1, 2, 3, and 4 mg of catalyst showed low quantitative E. coli growth inhibition, which was ~17% and 65%, 97%, 99%, respectively.[587]
Cu2O-g-C3N436 W fluorescent lamp with the cut-off filter of ~400 nm (λ > 400 nm)Bactericidal efficiency against B. subtilis, E. coli, S. aureus and P. aeruginosaThe maximum ZOI for Cu2O-g-C3N4 to B. subtilis E.coli, S. aureus and P. aeruginosa is 22 ± 1.67, 15 ± 1.08, 11 ± 1.22, 6 ± 0.09, respectively.[587]
g-C3N4/TiO2 KaoliniteXenon lamp with a 400 nm cut-off filterDisinfection ability towards S. aureusThe disinfection efficiency of g-C3N4/TiO2/kaolinite is ~4.3 log cfu/mL in 5 h.[588]
TiO2/g-C3N4500 W xenon arc lamp with a 400 nm cut-off filterAnti-fouling ability of E. coliTiO2/g-C3N4 showed an excellent E. coli removal with a permeate flux of 2 times higher than that of filtration alone.[589]
TiO2/g-C3N4-Bactericidal efficiency against E. coliThe bacterial survival rate for the TiO2 nanotube/g-C3N4 nanofilms is ~16%.[590]
g-C3N4/Ag-TiO2Xenon lampBactericidal efficiency against both Gram-negative 18 Escherichia-coli and Gram-positive Staphylococcus-aureusThe presence of Ag in g-C3N4/Ag-TiO2 structure could enhance water disinfection under visible light.[591]
g-C3N4/MoS2-Bi2O3300 W xenon arc lampBactericidal efficiency against E. colig-C3N4/Bi2O4 with the ratio of 1:0.5, could entirely inactivate 6-log10 cfu/mL E. coli.[586]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alaghmandfard, A.; Ghandi, K. A Comprehensive Review of Graphitic Carbon Nitride (g-C3N4)–Metal Oxide-Based Nanocomposites: Potential for Photocatalysis and Sensing. Nanomaterials 2022, 12, 294. https://doi.org/10.3390/nano12020294

AMA Style

Alaghmandfard A, Ghandi K. A Comprehensive Review of Graphitic Carbon Nitride (g-C3N4)–Metal Oxide-Based Nanocomposites: Potential for Photocatalysis and Sensing. Nanomaterials. 2022; 12(2):294. https://doi.org/10.3390/nano12020294

Chicago/Turabian Style

Alaghmandfard, Amirhossein, and Khashayar Ghandi. 2022. "A Comprehensive Review of Graphitic Carbon Nitride (g-C3N4)–Metal Oxide-Based Nanocomposites: Potential for Photocatalysis and Sensing" Nanomaterials 12, no. 2: 294. https://doi.org/10.3390/nano12020294

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