Next Article in Journal
Hydrotalcite-Supported Ag/Pd Bimetallic Nanoclusters Catalyzed Oxidation and One-Pot Aldol Reaction in Water
Previous Article in Journal
β-Cyclodextrin-Silica Hybrid: A Spatially Controllable Anchoring Strategy for Cu(II)/Cu(I) Complex Immobilization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Doping of Graphitic Carbon Nitride with Non-Metal Elements and Its Applications in Photocatalysis

1
Institute of Environmental Technology, VŠB-Technical University of Ostrava, 17. listopadu 15, 708 00 Ostrava-Poruba, Czech Republic
2
Department of Chemistry, VŠB-Technical University of Ostrava, 17. listopadu 15, 708 00 Ostrava-Poruba, Czech Republic
3
Chuiko Institute of Surface Chemistry of National Academy of Sciences of Ukraine, 17 General Naumov Street, 03164 Kyiv, Ukraine
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(10), 1119; https://doi.org/10.3390/catal10101119
Submission received: 31 August 2020 / Revised: 21 September 2020 / Accepted: 23 September 2020 / Published: 28 September 2020
(This article belongs to the Section Catalytic Materials)

Abstract

:
This review outlines the latest research into the design of graphitic carbon nitride (g-C3N4) with non-metal elements. The emphasis is put on modulation of composition and morphology of g-C3N4 doped with oxygen, sulfur, phosphor, nitrogen, carbon as well as nitrogen and carbon vacancies. Typically, the various methods of non-metal elements introducing in g-C3N4 have been explored to simultaneously tune the textural and electronic properties of g-C3N4 for improving its response to the entire visible light range, facilitating a charge separation, and prolonging a charge carrier lifetime. The application fields of such doped graphitic carbon nitride are summarized into three categories: CO2 reduction, H2-evolution, and organic contaminants degradation. This review shows some main directions and affords to design the g-C3N4 doping with non-metal elements for real photocatalytic applications.

1. Introduction

Graphitic carbon nitride has been of interest since Wang et al. published their paper about its ability of photocatalytic water splitting in 2009 [1]. It is a two-dimensional (2D) metal-free semiconducting material with the ability to absorb visible light due to the band gap energy of 2.7 eV. Other interesting properties, such as high thermal, physical, chemical, and photochemical stability, predetermines it for solar cell fabrications, imaging, sensing of some compounds, and also for photocatalysis [2,3,4,5,6,7,8,9,10]. On the other hand, there are some problems connected to fast recombination of photoinduced electrons and holes and low specific surface area, which must be solved.
Introducing non-metal elements into the g-C3N4 framework modulates its surface morphology, size of particles, electronic and optical properties, and other physico-chemical properties. Since bare g-C3N4 absorbs light up to 420 nm, the non-metal incorporation allows us to extend absorption of visible irradiation and to reduce recombination of photoinduced electrons and holes. Over the past few years, a lot of articles on non-metal doping have been published. This literature review is focused on recent progress in the synthesis and design of non-metal doped g-C3N4 for applications in photocatalysis.

2. Oxygen-Doped g-C3N4

The most common ways to introduce elemental oxygen into g-C3N4 (Table 1) are oxidation methods, such as oxidation of g-C3N4 with acids [11,12,13], thermal oxidation [14,15,16,17,18,19,20,21,22,23], oxidation with hydrogen peroxide H2O2 [24,25,26,27,28,29], hydrothermal treatment [30,31,32], pretreatment of synthetic precursors followed by thermal oxidation [33,34,35,36,37], and solvothermal methods [38,39,40].
Putri et al. performed the doping with oxygen atoms via hydrothermal treatment of bulk g-C3N4 with H2O2 at 120 °C [24]. The concentration of H2O2 influenced the specific surface area (SSA) of O-g-C3N4. The highest SSA of O-g-C3N4 was determined to be 85 m2·g−1 compared to 75 m2·g−1 for pristine g-C3N4. The photocatalytic experiment did not reveal the influence of SSA on activity of O-g-C3N4 photocatalysts. Fourier-transformed infrared spectroscopy (FTIR) allowed authors to get confirmation that the oxygen functionalization of g-C3N4 was achieved by its oxidation with H2O2. The new band at 1070 cm−1 was described as vibration of new stretching modes of N-O groups. The insertion of elemental oxygen into g-C3N4 was confirmed by X-ray photoelectron spectroscopy (XPS). The apparent new peak at the binding energy cca 530 eV was ascribed to oxygen doping into the g-C3N4 lattice. The peaks of adhering oxygen atoms directly bound to carbon atoms (C–O) and substitutional O in the form of N–C–O heterocyclic rings of the graphitic structure were observed on a wide scan survey. It was also confirmed that oxygen doping greatly influenced the carbon chemical state. An optical properties study demonstrated that oxygen doping caused band gap narrowing due to the formation of sub-gap impurity states that extend the light absorption of O-g-C3N4 (Figure 1).
The oxygen groups in the g-C3N4 lattice were supposed to form midgap states, that is, localized states in the band gap. It was observed the decreasing of the midgap energy in the values from 2.30 to 1.80 eV when 10 and 90 cm3 of H2O2 were used. At the same time, for the majority of O-g-C3N4 samples, the intrinsic band gap value was 2.85 eV. The band structure transformation with increasing oxygen doping level in g-C3N4 were suggested to occur according to the scheme presented in Figure 2.
Similar characteristics of O-g-C3N4 were reported by Zhang et al. [41], which performed the oxygen doping in the same way. The reducing of photogenerated carrier recombination was confirmed with photoluminescence (PL) method. The study of optical properties of O-g-C3N4 show that the H2O2 treatment had no effect on the band gap of the catalyst. It caused a significant effect on enlarging a photocatalytic response range of the obtained material and increased its absorption intensity.
Huang et al. performed the oxidative treatment of bulk g-C3N4 with peroxymonosulfate (PMS) at 60 °C [18]. As the SSA of g-C3N4 after the oxidation was approximately unchanged, it was suggested that PMS interacts with the surface groups of g-C3N4 without its exfoliation into a single or few layers. The atomic percentage of oxygen increased on O- g-C3N4 surface to 6.9% in comparison with 1.8% in g-C3N4 as confirmed by XPS. The detailed study of the chemical state of C, N, and O atoms in O-g-C3N4 allowed authors to report that the doped oxygen atoms mainly exist as carbonyl and carboxyl groups. The doping of more oxygen groups in O-g-C3N4 caused a decrease of the optical band gap from 2.82 to 2.79 eV. The doping of oxygen improved the separation of photo-generated carriers that was concluded from the decreasing PL intensity after oxidation of g-C3N4 with PMS.
Li et al. performed the synthesis of nanostructured O-g-C3N4 with C-O-C and C=O groups by a two-step thermal treatment process [19]. Increase of the O content was observed for every step of the treatment as was determined by elemental analysis (EA). The O content in pristine g-C3N4 and its products of the first and second steps of the treatment were 4.07%, 5.01%, and 8.43%, respectively. The decreasing of N content from 60.30% to 54.18% can be explained by substituting N atoms with O ones in the carbon nitride skeleton. The study of elemental valence state in the samples by XPS revealed that C-O-C groups were introduced into the precursor with copolymerization of urea and ammonium acetate at the first step of synthesis, and the C=O groups were introduced into O-g-C3N4 at the second thermal treatment step. The electron paramagnetic resonance (EPR) signal intensity study of O-g-C3N4 revealed internal electric field between O and tri-s-triazines that caused the remarkable enhancement of visible-light photocatalytic pollutant degradation. The presence of two types of oxygen functional groups caused the polarization of charge distribution, which formed an internal electric field. The negative pole played the oxygen-containing zone and the positive pole played the tri-s-triazine part.
The strategy of another group of researchers concerned the development of purposefully designing precursor for the further thermal treatment [42]. A new hydroxylated and carbonylated melamine for synthesis of porous O-g-C3N4 nanosheets (OCNNS) was obtained by a hydrothermally ethanol-assisted reforming of melamine at 200 °C. During the polymerization of the precursor, oxygen atoms of anchored hydroxyl groups contributed to oxygen doping, while alternative oxygen atoms in anchored carbonyl groups were simultaneously removed in the form of water forming porous nanostructures. The porous OCNNS demonstrated the increased specific surface area (67.4 m2·g−1) providing more active sites, and enhanced the transfer and separation of charges. The C–O bond formation in O-g-C3N4 was demonstrated with XPS. C–O peaks exhibited a progressively increased intensity with increasing amount of ethanol used for the precursor synthesis. The band gaps of O-g-C3N4 and g-C3N4 were determined to be 2.67 eV and 2.60 eV, respectively. It was established that O-g-C3N4 satisfied with the thermodynamic condition for photocatalytic H2 production, and its conduction band energy was upshifted, confirming the increasing of photoreduction capability.
The solvothermal method was applied by Wei et al. to fabricate a series of O-doped g-C3N4 photocatalyst [22]. The copolymerization of 1,3,5-trichlorotriazine and various amounts of dicyandiamide was performed at a relatively low temperature. The absorption band of ultraviolet-visible (UV-Vis) diffuse reflectance (DR) spectra of O-g-C3N4 was red-shifted, which evidenced that introducing oxygen extended the sensitivity of obtained materials to a wider range of light. Compared to pristine g-C3N4 with a band gap of 2.73 eV, O-g-C3N4 demonstrated narrowing of the band gap from 2.54 to 2.09 eV.

3. Nitrogen-Doped g-C3N4

The number of publications on obtaining nitrogen-doped graphitic carbon nitride (N-g-C3N4) is much lower in comparison with the number of publications on carbon, oxygen, phosphorus, and sulfur doped g-C3N4. The reason for the limited production of N-g-C3N4 is the application of non-environmentally friendly precursors, such as hydroxylammonium chloride and hydrazine hydrate. Some processes of the N-g-C3N4 synthesis generally contain complicated and long-lasting steps. The presented reports on the synthesis and characterization of N-g-C3N4 and its application in photocatalysis inspired researchers to develop appropriate methods of introducing nitrogen into the g-C3N4 lattice (Table 2). Different reaction paths were applied to increase the N content in g-C3N4 by introducing nitrogen atoms into g-C3N4 matrices [44,45,46,47,48,49,50,51,52,53,54,55] and by the modification of g-C3N4 surface [56,57,58,59,60,61].
An extremely rapid method of the high-nitrogen content carbon nitride production was offered by Miller et al. [44,45,46]. They decomposed a single molecular precursor, trichloromelamine, which briefly generated internal temperatures near 400 °C and resulted in amorphous C3N4+x, where 0.5 < x < 0.8. According to the report the carbon centers of the obtained materials primarily had sp2 hybridization, the layered structure consisted of triazine (C3N3) rings, and nitrogen species bound the triazines. The N-g-C3N4 was thermally stable up to 600 °C, possessed a moderately porous structure with nanospherical morphologies, and demonstrated a blue photoluminescence.
The various nitrogen-rich compounds were used as the precursors of N-g-C3N4. Huynh et al. synthesized 3,6-di(azido)-1,2,4,5-tetrazine for obtaining nitrogen-rich carbon nitrides C3N4 and C3N5 [47]. The high-nitrogen contained nitrides were also synthesized by thermal decomposition of 2,4,6-triazido-1,3,5-triazine [48], 2,5,8 triazido-s-heptazine [44,45,46]. However, these types of precursors are often thermodynamically unstable, shock and impact sensitive, and should be handled with caution.
Several groups reported on obtaining the nitrogen doped graphitic carbon nitride (C3N4+x) [49,50,51,52,62]. Fang et al. performed the thermal condensation of precursor with hydrazine hydrate [49]. They confirmed that the nitrogen atom substitutes the sp2 carbon atom in the resultant N-g-C3N4. According to EA the N/C mass ratio was 1.68 for C3N4+x, that is, higher than that (1.60) of pristine g-C3N4.
N-g-C3N4 with the high SSA was obtained by Xu et al. via secondary calcinations of g-C3N4 at different temperatures [50]. The highest nitrogen doping ratio and the largest SSA was observed for g-C3N4 that was post treated at 590 °C, which was confirmed with XPS, X-ray diffraction (XRD), and energy dispersion spectroscopy (EDS) methods. The XPS analysis allowed authors to conclude that the sp2 C atom was replaced by the nitrogen atom. According to the EDS analysis, the highest value of N content in the N-g-C3N4 was 57.57 at.% and the lowest for undoped g-C3N4 was 56.88 at.%. The secondary calcination at 590 °C caused the SSA to increase of about 27 times (from 4.62 to 128.06 m2·g−1 for non-treated and post thermal treated g-C3N4, respectively). The enhanced visible light absorption capacity was confirmed with UV-Vis DR spectra, which demonstrated the absorption edge red shift from 464 nm to 515 nm with the secondary calcination temperature increasing. The authors suggested that the shift was caused by the change of electronic structure due to the nitrogen doping. The bandgap of g-C3N4 narrowed from 2.67 to 2.41 eV after the secondary calcination step.
Jiang et al. used the same precursor as Xu et al. [50] for obtaining of g-C3N4, which they modified with N-N dimethylformamide (DMF) and then treated at 550 °C for 4 h under static air [52]. The SSA of g-C3N4 modified with DMF (g-C3N4/DMF) was 42.18 m2·g−1 in comparison with 18.36 m2·g−1 for non-treated g-C3N4. The further increasing of the specific surface area to 74.79 m2·g−1 was reached after the thermal treatment of g-C3N4 with DMF for 4 h. The porous nanosheet architecture of N-g-C3N4 caused the increasing of SSA confirmed with transmission electron microscopy (TEM) and atomic force microscopy (AFM). The C/N molar ratios were 0.692, 0.679, and 0.674 for the pristine g-C3N4, g-C3N4/DMF, and exfoliated N-g-C3N4, respectively. The authors proposed the copolymerization route of dicyandiamide and DMF. According to the XPS analysis the substitution of carbon with nitrogen was concluded to occur. The quantitative XPS analysis showed that the C/N molar ratio of N-g-C3N4 was 0.67, and for pristine it was 0.76. The N doping and nanosheet construction of g-C3N4 expanded an absorption band of obtained materials and enhanced its ability to harvest visible light. The band gaps of pristine and N-g-C3N4 were estimated to be 2.51 eV and 2.54 eV, respectively. The transfer and separation efficiency of charge carries in N-g-C3N4 materials were studied with PL spectroscopy. The decreasing of PL intensity of N-g-C3N4 in comparison with g-C3N4 was explained by the presence of midgap states produced by the N doping. They served as separation centers to capture photoexcited electrons averting recombination processes.
Guo et al. synthesized N-doped porous g-C3N4 with enhanced ability of photocatalytic H2 production through the one-step thermal copolymerization of urea and DMF [51]. No influence of DMF on the original graphitic C-N network was revealed with FTIR spectroscopy. According to XPS and EA g-C3N4, that which was obtained by the copolymerization of urea and DMF, was enriched with nitrogen. The results of 13C nuclear magnetic resonance (NMR) and Raman analysis allowed the authors to conclude that the replacement of C atoms in triazine with N atoms occurred. The modification with DMF promoted the increasing of light absorption in visible region. The obvious red-shift of the absorption edge of N-g-C3N4 was suggested to be the result of incorporated N atoms that caused increased delocalization of π-electrons. The analysis of UV-Vis DR spectra of N-g-C3N4 revealed the narrowing of band gap to 2.69 eV compared to 2.75 eV of pristine g-C3N4.
The supramolecular self-assembly strategy for obtaining of porous N-g-C3N4 nanotubes with the high adsorption of CO2 was offered by Mo et al. [62]. In the first step, melamine and hydroxylammonium chloride were used to obtain a supramolecular intermediate at 120 °C (Figure 3a). Secondly, the supramolecular intermediate was heated at 520 °C under NH3, air, Ar, and N2 atmospheres. From the scanning electron microscopy (SEM) analysis of N-g-C3N4 obtained under the thermal treatment in various gases, the authors revealed the porous structure only in g-C3N4 prepared under the NH3 atmosphere (Figure 3b–d). The sharp increase of SSA was observed for the N-g-C3N4 materials. The determined values of the SSA of bulk g-C3N4 and materials, which were obtained by treatment in Ar, air, N2, and NH3, were 8.6, 56.9, 78.6, 82.3, and 108.9 m2·g−1, respectively. The highest content of amino groups grafted into g-C3N4 nanotubes during the time of the supramolecular intermediate polymerization process was observed for the porous materials treated under NH3 (g-C3N4/NH3). The EA results indicated a significantly lower C/N atomic ratio of g-C3N4/NH3 than those of g-C3N4 samples synthesized under other atmospheres. The determined C/N atomic ratio for the pristine g-C3N4 and materials obtained in NH3, Ar, and N2, air were 0.69, 0.55, 0.59, 0.58, and 0.58 m2·g−1, respectively. Based on the XPS analysis, the authors concluded that in the NH3 atmosphere, amino defects corresponding to terminal isolated amino groups were produced as a result of the opening of C-N bonds. The influence of amino groups on the adsorption of CO2 and desorption of CO was studied with density functional theory (DFT) calculations. The DFT calculations demonstrated that g-C3N4/NH3 not only increased the CO2 adsorption energy, but also decreased the CO desorption energy that caused the higher photocatalytic CO2 reduction activity.
Hao at al. incorporated nitrogen atoms into g-C3N4 by the hydrothermal treatment of exfoliated material using DMF [57] according to the scheme presented in Figure 4. The X-ray diffraction (XRD) and FTIR analysis of the obtained 2D g-C3N4 nanosheets revealed that the basic characteristics of g-C3N4 structure were not influenced by the surface nitrogen modification. Formation of new C-N bonds due to N atoms insertion into the g-C3N4 lattice by filling of removed oxygen atom seats was confirmed by XPS. The authors distinguished three types of the sp2 hybridized nitrogen: Pyridinic nitrogen (C-N=C), tertiary nitrogen (N-C3), and noncondensing amino group (C-N-H), respectively. They confirmed the highest content of pyridinic N atoms in N-g-C3N4 that is beneficial to the photocatalytic production of hydrogen. According to XPS, no new N-O bonds were formed. The EDS elemental mapping allowed the authors to evidence the increased nitrogen content in the surface of N-g-C3N4. The nitrogen-doped g-C3N4 demonstrated enhanced visible absorption ability and the increasing lifetime of its excited state.
Amine-functionalized g-C3N4 was obtained by its treatment with monoethanolamine (MEA) [58]. The ratio of amino groups in the treated sample was shown to be obviously higher due to the introduction of MEA molecules. Nearly the same absorption edges of amine-functionalized g-C3N4 evidenced the absence of any doping effect and thus changes in the intrinsic electronic structure of g-C3N4. The amine-functionalized materials demonstrated the increased adsorption of CO2 caused by the chemical interaction between CO2 and amino groups. The treatment of g-C3N4 with MEA resulted in the stronger adsorption of hydroxyl ions caused decreasing of zeta potentials in aqueous solutions. The zeta potential at pH = 7 of the amine-functionalized and pristine g-C3N4 were −19.44 mV and −31.67 mV, respectively.
Yan et al. applied a NH4Cl-assisted chemical blowing method to prepare nitrogen-rich graphitic carbon nitride nanosheets [59]. They revealed that under the thermal treatment of a NH4Cl and melamine mixture, generated gases detached g-C3N4 sheets from each other and decreased the polymerization of carbon nitride precursor subsequently forming nitrogen-doped nanosheets. The content of nitrogen in the obtained materials was higher (47.32 at.%) in comparison with pristine g-C3N4 (45.59 at.%). The band gaps calculated from UV-Vis DR spectra increased from 2.73 eV for g-C3N4 to 2.79 eV for N-g-C3N4. It was shown that the photoluminescence intensity of these materials was diminished when the morphology changed from bulky to nanosheet-like, due to suppression of electron–hole recombination in nanosheets.
Zhou et al. synthesized N-g-C3N4 by the hydrothermal treatments of urea with small amount of citric acid [63]. The authors suggested that citric acid was a carbon source and promoted the formation of N-g-C3N4 by the reaction with urea. Changes in the structure, SSA, and elemental compositions of obtained N-g-C3N4 were not observed. N-g-C3N4 demonstrated a little red shift of the intrinsic absorption edge in comparison with that of g-C3N4. The color of the materials changed from light to brown yellow with increase of the citric acid content in the reaction mixture. The authors suggested a lone pair electron on g-C3N4 may trigger π-electron delocalization in this conjugated system. This material also demonstrated fluorescence quenching that evidenced the suppressed recombination of photogenerated charge carriers.
Wang et al. tried to dope g-C3N4 with nitrogen by the thermal polymerization of urea under the nitrogen atmosphere [53]. The obtained N-g-C3N4 demonstrated the decreasing C/N ratio 0.71 compared to 0.73 of g-C3N4 as was revealed by the XPS analysis. The authors suggested that the decreasing of C/N ratio in N-g-C3N4 occurred due to the formation of carbon vacancies that promoted the formation of tri-s-triazine subunits. The intrinsic bandgaps of N-g-C3N4 obtained in the nitrogen atmosphere increased to 2.70 eV from 2.58 eV of g-C3N4 obtained in air.
Tian et al. discovered that the pre-hydrothermal treatment of urea and melamine mixture (molar ratio of urea:melamine = 3:1) at 180 °C promoted an irreversible monoclinic to orthorhombic melamine phase transformation, calcination of which caused the formation of mesoporous g-C3N4 with enhanced photocatalytic activity [64]. The formation of ultrathin N-g-C3N4 nanosheets with a thickness of ~3 nm that consisted of 7 or 8 at. layers were confirmed with AFM. The SSA of N-g-C3N4 was 39.1 m2·g−1 with an average pore diameter of 46.1 nm, which was 9 times higher than pristine g-C3N4. The introduction of N atoms into the g-C3N4 matrices was confirmed with the EA and XPS analyses. The band gap of N-g-C3N4 decreased to 2.47 eV in comparison with pristine g-C3N4 (2.70 eV). The multiple reflection of incident light across the porous open network structure of the N-rich materials caused their stronger light harvesting capability that appeared as enhanced photo-absorption below 450 nm.
Table 2. Synthetic methods, applications, and photocatalytic efficiency of N-doped g-C3N4.
Table 2. Synthetic methods, applications, and photocatalytic efficiency of N-doped g-C3N4.
PrecursorSynthetic MethodC/N Atomic Ratio,
Doped (Pristine)
Photocatalytic ProcessConditions of the ProcessEfficiency Doped/PristineReferences
Melamine, N-N dimethylformamide (DMF)Hydrothermal treatment0.407 (0.540) by EDS,
0.457 (0.575)
by Organic Elemental Analysis (OEA)
H2
evolution
300W Xe lamp, filter λ > 400 nm), co-catalysts Pt nanoparticles, triethanolamine (TEOA) as a hole quencher128.5 h−1/
58.6 h−1,
[57]
Melamine, hydrazine hydrateThermal condensation0.67 (0.73) by elemental analysisH2 evolution300W Xe lamp, filter λ > 400 nm, TEOA (10 vol.%, Pt co-catalyst (3 wt.%)44.28 μmol·h−1/
7.86 μmol⋅h−1
[49]
dicyandiamideSecondary calcination57.57 (56.88) at.% by EDSmethylene blue
degradation
300W Xe lamp, filter λ > 420 nm0.02355 min−1/
0.00829 min−1
[50]
Dicyandiamide, DMFThermal copolymerization0.67 (0.76) by XPSTetracycline (TC) degradation300W Xe lamp, filter λ > 420 nm76.78 (52.21)% of TC was degraded in 60 min[52]
Urea, DMFThermal copolymerization0.74 (0.59) at.% by organic elemental analysisH2 evolution300W Xe lamp, filter λ > 400 nm, TEOA (10 vol.%), Pt co-catalyst (3 wt.%)5268 μmol g−1·h−1/
3579 μmol g−1·h−1/
[51]
Urea, monoethanolamineAmine functionalization of g-C3N4CO2 reduction300W Xe lamp, gas phase reactionCH4: 0.34 μmol·h−1·g−1/trace;
CH3OH: 0.26 μmol·h−1·g−1/
0.26 μmol·h−1·g−1
[58]
Melamine,
NH4Cl
Thermal polymerizationN content (at.%)
doped 47.32 doped;
45.59 pristine
RhB degradation 300W Xe lamp, 420 nm cutoff filter,0.01954 min−1/
0.00391 min−1
[59]
Melamine,
NH4Cl
Second-calcination approach1.14 (1.63) by XPS analysisH2 evolution420-nm LED, lactic acid (10 vol%) solution, (Pt co-catalyst (1 wt%)15.5 mmol·h−1/
7.5 mmol·h−1
[65]
Melamine,
NH4Cl
Precursor formation by hydrothermal method; thermal polymerization in NH3, N2, Ar, air0.55-0.59 (0.69) by elemental analysisCO2 reduction 300 W Xe-lamp, CoCl2, 2,2-bipyridine, TEOA and methyl cyanide 103.6 μmol·g−1·h−1/
6.1 μmol·g−1·h−1
[62]
Citric acid, ureaThermal polymerizationwithout changeH2 evolution300W Xe lamp, 420 nm cutoff filter, Pt nanoparticles (3 wt.%), triethanolamine as a hole quencher64 μmol·h−1/15 μmol·h−1[63]
Dicyandiamide, citric acid, ureaThermal polymerizationIndomethacine degradation350 W xenon lamp with 420 nm cutoff filter; 500 W mercury lamp; 350 W xenon lamp with a 290 nm cut-off filterPhotocatalytic activity was 1.1 (UV light irradiation), 1.8 (simulated sunlight),
and 13.6 (visible light irradiation) times higher than that of pristine g-C3N4
[66]
Melamine, ureaHydrothermal treatmentC:N mass ratio doped 0.53/0.73 by OEA;
C:N atomic ratio 0.47/0.72 by quantitative XPS
H2 evolution300W Xe lamp, 420 nm cutoff filter, Pt nanoparticles (1 wt.%), 20% vol. lactic acid3579 μmol·h−1·g−1/
147 μmol·h−1·g−1
[64]
Urea, N2Thermal polymerizationdoped 0.71/0.73 by XPSBisphenol A oxidation; Cr(VI) reduction300W Xe lamp, 420 nm cutoff filterComplete degradation of BPA 60 min/90 min;
Photoreduction of Cr(VI) over 120 min: 10%/60%
[34]

4. Carbon-Doped g-C3N4

It has been proposed that g-C3N4-carbonaceous compound hybrids can enhance visible-light absorption and improve photoinduced charge separation efficiency. The various allotropies and hetero-structures of carbon were used for incorporation into g-C3N4 to optimize its conductivity and photocatalytic activity [67,68,69,70,71,72,73,74,75,76,77,78,79] (Table 3). The workable and effective way to design the surface structures and electronic properties of g-C3N4 can also be provided by the self-doping of g-C3N4 [80,81,82,83,84,85,86,87].
The carbon doped g-C3N4 (C-g-C3N4) composite was fabricated by Cao et al. in a facile one-pot way by the calcination of dicyanamide and small amounts of dimethylformamide as a cost-effective carbon source [71]. The increasing C/N mass ratio determined by EA of 0.61 for C-g-C3N4 in comparison with 0.59 for pristine g-C3N4 evidenced successful carbon incorporation. The substitution of C atoms on original sites of bridging N was confirmed by the XPS study. The authors found that the content of bridging N in C-g-C3N4 decreased from 17.28% to 16.95% due to replacement of bridging nitrogen by carbon atoms. C-g-C3N4 contained the higher content of C–N=C groups (71.17%) compared with pristine g-C3N4 (70.75%), which benefited the formation of π-conjugated system. The C-doping caused a minor increasing of SSA from 16.63 m2·g−1 for g-C3N4 to 17.80 m2·g−1 for C doped g-C3N4. Typical semiconductor intrinsic absorption in UV range was observed for both g-C3N4. At the same time, the edge of C-g-C3N4 absorption band demonstrated some red shifts with respect to the pristine material. Additionally, remarkable enhancing of the absorption intensity of C-g-C3N4 in visible-light region was observed. The values of band gap of pristine g-C3N4 and C-g-C3N4 calculated by Tauc plots were 2.72 eV and 2.66 eV, respectively. The values of conduction band (CB) determined with the photoelectrochemical method were −1.21 V and −0.94 V (vs. Ag/AgCl electrode) for pristine g-C3N4 and C-g-C3N4, respectively. The valence band (VB) of C-doped g-C3N4 1.92 eV was shifted on 0.21 eV with respect to g-C3N4 (1.71 eV). The influence of C-doping on PL of g-C3N4, which was originated from recombination of holes at valence band and electrons at conduction band, was studied. C-g-C3N4 demonstrated the higher intensity of PL emission with respect to the pristine one. The authors concluded that C-g-C3N4 behaves as an electron buffer that effectively speeds up excited electrons and delocalization of π-electrons in the conjugated system causing the separation ability of photo-excited holes and electrons. C-g-C3N4 demonstrated 4.3 times higher photocurrent density compared with pristine g-C3N4 under visible light.
A facile one-step thermal condensation method using an agar melamine gel (AMG) as the precursor was offered by Wang et al. to synthesize mesoporous C-g-C3N4 ultrathin nanosheets (C/CNNS) [79]. The agar was applied as a soft template, which was carbonized between g-C3N4 layers during the thermal polymerization. The ultrathin nanosheet structures and the typical wrinkled morphology was found on TEM images. The AFM analysis evidenced that C/CNNS consisted of approximately 10 single layers of g-C3N4. The absorption edge of C/CNNS was shifted to the longer wavelengths with the increasing carbon content. Similarly, the color of the C/CNNS nanosheets was changed from light yellow to dark grey (Figure 5a). The values of band gap energies determined by the Tauc method were 2.62 eV and 2.56 eV for bulk g-C3N4 and C/CNNS, respectively (Figure 5b). The CB edge potential −0.61 eV of C/CNNS was negatively shifted in relation to the CB edge potential −0.42 eV of g-C3N4 (Figure 5c,d). The observed decreasing of PL intensity for C/CNNS compared to pristine g-C3N4 evidenced the suppressing of electron–hole pairs recombination induced by the carbon doping.
Ran et al. performed a simple thermal polymerization method to obtain C-g-C3N4 photocatalysts with high NO removal efficiency under visible light [82]. C-g-C3N4 prepared by the co-pyrolysis of urea and saccharose possessed high SSA: 81 and 118 m2·g−1, for bulk and C-g-C3N4, respectively. It improved the photocatalytic performance due to the formation of more active sites to assist the transfer of reactants. The substitution of N with C favored charge transportation and was observed by room temperature, solid-state electron paramagnetic resonance (EPR). The PL intensity greatly decreased with the increasing of carbon amount confirming that the carbon substitution suppressed photogenerated electron–hole recombination. The C-g-C3N4 samples demonstrated improved visible-light absorption and narrowed band gap in comparison with pristine g-C3N4. The values of band gap were 2.71 eV for pristine and 2.65 eV C-g-C3N4, respectively. The authors demonstrated that C-g-C3N4 promoted the reactants activation, improved photo-generated charges separation, and assisted generation of radicals for the NO oxidation under visible light.
Long et al. obtained C60/g-C3N4 nanowire composites for the photocatalytic H2 evolution process by the thermal treatment of urea and C60 nanorods mixture [76]. The insignificant increasing of SSA from 104.77 m2·g−1 (for pristine g-C3N4) to 117.47 m2·g−1 was observed. Both samples were mesoporous with the main pore width of 2-6 nm. The incorporation of C60 nanorods in g-C3N4 improved visible light absorbance. The band gap of the C60/g-C3N4 photocatalyst was narrower (2.57 eV) compared to pristine g-C3N4 (2.75 eV). The improved ability of C60/g-C3N4 to separate photogenerated electron–hole pairs was concluded from g-C3N4 fluorescence quenching due to modification with C60. The C60/g-C3N4 nanocomposite produced more photogenerated electrons as demonstrated by the highest transient photocurrent responses compared to pristine g-C3N4. The improvement of C60/g-C3N4 photocatalytic H2 production was explained by synergy of g-C3N4 and C60, which allowed the photogenerated charges separation.
The combination of carbon dots (CDs) with g-C3N4 (CDs/g-C3N4) activated by hydrogen peroxide allowed Asadzadeh et al. to obtain photocatalysts with exceptional activity upon visible-light irradiation [67]. The activation of g-C3N4 with H2O2 increased its SSA from 14.6 to 45.1 m2·g−1 due to exfoliation during the treatment process. However, the adhering of CDs blocked micropores of the treated g-C3N4 that caused the SSA decrease to 19.7 m2·g−1. The homogeneous dispersion of CDs on the surface of activated g-C3N4 was confirmed by the EDS analysis. The EDS elemental mappings of CDs/g-C3N4 demonstrated the uniform distribution of C, N, and O elements evidencing the binary nanocomposite formation. The presence of two components in CDs/g-C3N4 was also confirmed by high-resolution TEM (HRTEM) images. The d-spacing of g-C3N4 was determined to be 0.326 nm, corresponding to the lattice fringe of (002) planes, while the spacing of 0.320 nm corresponded to the (002) plane of CDs. The absorption band of the pristine g-C3N4 edge at 460 nm and was red shifted to 475 nm in the spectrum of the H2O2 activated g-C3N4. The absorbance of CDs/g-C3N4 demonstrated a slight shift to the visible region. The as-synthesized CDs demonstrated PL emissions located at 504 nm that shifted when excited by various wavelength lights (from 500 to 900 nm). The authors suggested a multiphoton active process took place, in which absorption of two or more photons occurred simultaneously and caused emission of light at a shorter wavelength compared to the excitation wavelength. The PL study of CDs/g-C3N4 was not reported. The thermal decomposition of the CDs/g-C3N4 occurred at lower temperatures than that of pristine g-C3N4 due to the promoted exfoliation and integrated CDs.

5. Sulfur-Doped g-C3N4

The doping of g-C3N4 with sulfur (S-g-C3N4) caused changes in its electronic structure, adjustment to the position of CB and VB, carrier mobility enhancing, and as a result, improvement of photocatalytic activity (Table 4). The incorporation of S into g-C3N4 was performed by the thermal treatment under H2S [89], the thermal polymerization of S-containing precursors [90,91,92,93], the thermal copolymerization with S-containing compounds [94,95,96,97,98,99,100,101,102,103,104,105], by a gas-templating method [106], and by sulfuring and treatment of g-C3N4 [107].
Li et al. applied thioacetamide (TAA) as a sulfur source for obtaining S-doped terminal-methylated g-C3N4 (SM-g-C3N4) nanosheets by a one-pot copolymerization process [103]. The thioacetamide also performed a blocking function during the polymerization to generate structure edge defects, which caused enlargement of SSA. The highest SSA of SM-C3N4 was 90.5 m2·g−1 compared to 42.6 m2·g−1 of pristine g-C3N4. The XPS analysis revealed the increased C/N ratio in SM-g-C3N4 (1.26) compared to g-C3N4 (0.81) that the authors associated with methyl groups introducing. The increased number of the C-NHx (x = 1, 2) groups that were not involved in the polymerization due to a blocking effect of terminal methyl was confirmed with XPS. The remarkable shift of absorption edge to near-infrared region was observed in the UV-vis DR spectra of SM-g-C3N4 samples with the increasing amount of TAA used for sulfur incorporation. The band gaps of SM-g-C3N4 nanosheets decreased from 2.62 (for g-C3N4) to 1.85 eV (the optimum photocatalyst). The doping of sulfur into methylated melon units was confirmed with DFT calculation to promote the valence band splitting near the Fermi level and caused a midgap electronic state formation. As a result, a significant decrease of bandgap about 0.7 eV occurred. The band structures of the SM-g-C3N4 nanosheets were calculated based on the band gaps obtained from UV-vis DRS. The enhancement of non-radiative recombination rates after introducing sulfur and terminal methyl groups in melon was concluded from the PL study. From DFT calculations it was concluded that an internal electric field was formed due to changes in local charges allocation and lattices strain in melon units of SM-g-C3N4 nanosheets and thus the separation of electron–hole pairs was improved.
Lv et al. combined doping of sulfur into g-C3N4 in situ and thermal oxidative etching treatment of obtained S-g-C3N4 [90]. The thermal oxidative etching caused the breaking of H-bonds between layers of g-C3N4 nanosheets and decreased their thickness to 4.0 nm. S-g-C3N4 nanosheets treated over 3 h demonstrated increased SSA to 226.9 m2·g−1 compared to 16.6 m2·g−1 for pristine g-C3N4. An apparent blue shift of absorption band edge from 470 to 420 nm was demonstrated compared to pristine g-C3N4. However, in comparison with the analog without sulfur, the absorption band edge of S-g-C3N4 nanosheets was red-shifted by 5-10 nm. The band gap of pristine g-C3N4 and the nanosheets with and without S-doping was calculated to be 2.28, 2.73, and 2.85 eV. The mass percent of the S atom in the samples estimated by XPS demonstrated that thermal oxidative etching treatment caused the increased S content in g-C3N4. The content of S in the S-g-C3N4 nanosheets and S-g-C3N4 was calculated to be 1.58 and 0.51 wt.%, respectively. The analysis of the O/C and N/C ratios in the S-g-C3N4 before and after oxidative etching treatment revealed the formation of more surface O species and surface N defects after the etching treatment. It was also suggested that the doping of sulfur was favorable to the formation of surface N defects and O species. The S-g-C3N4 nanosheets demonstrated highest capability for the separation of photogenerated charge carrier and this material was stable for up to 36 h during the photocatalytic H2 evolution.
Ke et al. synthesized the graphene-like S-g-C3N4 by the thermal treatment of a urea and benzyl disulfide mixture at 560–650 °C in the Ar flow [101]. The pronounced increase of SSA was observed from 20.2 m2·g−1 of pristine g-C3N4 to 298.2 m2·g−1 of S-g-C3N4. The XPS analysis allowed the authors to evidence the formation of C–S bonding by replacing the latticed N with S, and the presence of C–SOx–C bond. The transformation of g-C3N4 to graphene-like S-g-C3N4 was achieved by weakening planar H-bonding when S partially replaced N. An obvious red shift of the absorption edges of S-g-C3N4 to 530 nm compared with that of g-C3N4 at 458 nm was observed (Figure 6a). Accordingly, the band gap energy decreased from 2.74 eV (g-C3N4) to 2.10 eV (S-g-C3N4) (Figure 6b). The calculation of band potentials of the samples vs. standard hydrogen electrode evidenced the formation of electronic structures with the elevated CB and VB potentials in S-g-C3N4 (Figure 6c). The as-designed S-g-C3N4 demonstrated the high photocatalytic elimination efficiency of UO22+ under visible-light illumination. The narrowed band gap with upshifting of CB and VB potentials, and the excellent efficiency of charge transfer and carrier utilization, caused the remarkable photoactivity of S-g-C3N4.
Fan et al. constructed S-g-C3N4 rods with increased SSA and tuned their band gap to enhance photocatalytic activity [95]. They conducted the pyrolysis of supramolecular melamine-trithiocyanuric acid complexes (MT) at 500–650 °C (Figure 7). S-g-C3N4 demonstrated the increased SSA of 52 m2·g−1 compared to pristine g-C3N4 (15 m2·g−1). The composition study by the elemental analysis showed similar atomic C/N ratios (0.67–0.69) of S-g-C3N4 with that of pristine g-C3N4 (0.65). The presence of about 2% of hydrogen in pristine g-C3N4 was attributed to the un-condensed amino groups. With the increasing temperature of the treatment, the H content decreased due to the higher condensation yield. The content of S in S-g-C3N4 decreased from 0.63 to 0.27 wt.% with the increasing synthesis temperature. The sulfur oxide species formed during the calcination were found. The improved optical absorption in the range of 450–600 nm was observed for all S-g-C3N4 rods. The corresponding band gaps estimated from the absorption spectra decreased after the S-doping from 2.70 eV of pristine g-C3N4 to 2.56 eV of S-g-C3N4. The decreasing of fluorescence intensity was observed for S-g-C3N4 evidencing enhancement of the separation efficiency of photogenerated electron−hole pairs. The S-g-C3N4 rods demonstrated the high adsorption and photocatalytic activity on the RhB decomposition under visible light.

6. P-Doped g-C3N4

To overcome the inherent drawbacks of g-C3N4 and to promote its photocatalytic activity, the doping of phosphorus into g-C3N4 (P-g-C3N4) was performed recently by many researchers [108,109,110,111,112,113,114,115,116,117,118,119,120,121]. The band gap structure regulation and the improvement of carrier separation efficiency were reached (Table 5). The most commonly used method for obtaining of P-g-C3N4 was the thermal modification of pristine g-C3N4 with sources of P [109,115,117] and their thermal condensation [108,110,118,119,122,123,124].
Su et al. performed the thermal condensation of adenosine phosphate and urea followed by thermal exfoliation to obtain porous P-g-C3N4 nanosheets [118]. The increasing of SSA to 84.8 m2·g−1 was reached compared to 66.6 m2·g−1 of pristine g-C3N4. The XPS analysis indicated the presence of P in the nanosheets in terms of P=N and P–N bonds. According to EA the content of P was 2.17 atomic%. The lower C/N ratio of 0.71 of P-g-C3N4 compared to 0.75 of pristine g-C3N4 was observed. The increasing of the O atomic% from 2.8% of pristine to 7.65% of the P-g-C3N4 nanosheets was ascribed to the higher content of C–OH bonds in the edge of g-C3N4 as a result of the thermal exfoliation. The authors suggested the replacement of C atoms with P ones. The lowest photogenerated electron–hole pairs recombination rate of the nanosheets was revealed by PL study. The midgap states and the expansion of π-electron conjugated system by P atom introduction in g-C3N4 was suggested due to the red shift of emission peaks about 70 nm compared to pristine g-C3N4. The more effective photoinduced carrier separation was confirmed by the fluorescence decay study and the quantitative confirmation of PL quenching was obtained. The P-g-C3N4 nanosheets demonstrated significant enhancement of visible-light absorption in the region of 450-750 nm. The nanosheets demonstrated a distinct Urbach tail absorption band, which the authors attributed to the n→π* electron transition from lone electron pairs of edge nitrogen atoms in the heptazine parts due to the formation of incompletely symmetric planar modes. The transition energy was estimated at 1.91 eV based on the Urbach tail absorption, which is associated with midgap states. The calculated band gaps of undoped and P-g-C3N4 was 2.73 and 2.56 eV, respectively. It was shown that the P-doping changed the original band gap structure due to the formation of impurity level between VB and CB. Decrease in the VB position from 1.7 to 1.57 eV was suggested as a result of defects generation under the thermal exfoliation. The greatly improved photocatalytic H2 evolution was reached using the with P-g-C3N4 nanosheets.
The nano-structuring and P-doping strategy was implemented by Zhao et al. via a one-pot process for improvement of the g-C3N4 photocatalytic hydrogen evolution activity under visible light [121]. The aqueous solution of dicyandiamide, NH4Cl (gas-template), and (NH4)2HPO4 (P dopant) was freeze-dried and the obtained material was calcined at 550 °C over 4 h. The SSA of P-g-C3N4 was determined to be 36.4 m2·g−1, which is higher than 5.5 m2·g−1 of pristine g-C3N4. The authors confirmed that the successful nano-structuring of g-C3N4 was due to the gas-template application. The XPS analysis evidenced the successful doping of P element in g-C3N4 and the probable replacement of C with P forming P-N bonds. P-g-C3N4 contained 2.05 wt.% of P, which was determined by XPS.
In contradiction to the previous study, Su et al. [118] found the incorporation of P caused the blue shift of intrinsic absorption edge of g-C3N4 from 458 nm to 440 nm. Accordingly, the band gap was changed from 2.74 eV to 2.90 eV. The band gap enlargement was explained by the quantum size effect. The PL study revealed that the recombination of photo-induced charge carriers in g-C3N4 was efficiently diminished due to the P-doping and the nano-structure formation. The photocurrent density test was applied to confirm the efficiency of charge separation and migration in P-g-C3N4. The photocurrent density of P-g-C3N4 was 0.51 μmol·g−1·h−1·μA−1 compared to 0.06 μmol·g−1·h−1·μA−1 of pristine g-C3N4.
P-doped tubular g-C3N4 (PT-g-C3N4) with surface defects was obtained by Guo et al. through a phosphorus-contained compounds-assisted hydrothermal method [125]. The melamine was used as a raw material. Sodium pyrophosphate, ammonium phosphate, sodium hypophosphite, and sodium phosphite were used as P-containing compounds. A hexagonal tube-shaped nanostructure of PT-g-C3N4 demonstrated the increasing of SSA from 8.6 to 24.5-32.4 m2·g−1 of pristine g-C3N4 and PT-g-C3N4 from various P sources, respectively. According to the XPS analysis, the replacement of C in the triazine moieties with P to form P-N bonds during synthesis was suggested. The content of P in PT-g-C3N4 varied from 0.32 to 0.87 wt.%. The highest P ratio was reached using sodium pyrophosphate. The XPS study also revealed decreasing of the C/N surface atom ratio from 0.74 (PT-g-C3N4) to 0.69 (g-C3N4) due to carbon defects formation during the synthesis. It was also established that carbon defects exist only on the surface of g-C3N4 as the C/N mass ratio determined for pristine and PT-g-C3N4 samples demonstrated the same value of 0.68. The P-doping and the tube structure expanded the visible-light absorption due to numerous reflections of light within the PT-g-C3N4 tubes. The band gap of PT-g-C3N4 obtained from ammonium phosphate was calculated at 2.63 eV compared to 2.75 eV of g-C3N4. The PL and photocurrent measurements confirmed the decrease of photo-induced electrons and holes recombination rate. A low resistance for interfacial charge transfer from the g-C3N4 photocatalysts to reacting molecules was observed for the PT-g-C3N4 samples using electrochemical impedance spectroscopy (EIS).
Zhou et al. reported on the synthesis of P-g-C3N4 by the thermal copolymerization of hexachlorocyclotriphosphazene and guanidinium hydrochloride, a non-expensive and environmentally friendly compound [126]. The obtained P-doped photocatalysts showed the high photocatalytic performance both in the H2 evolution and the photodecomposition of organic dyes. The synthesized P-g-C3N4 demonstrated the increased SSA of 40.5 m2·g−1 compared to 26.86 m2·g−1 of pristine g-C3N4. The optical properties study revealed that the P-g-C3N4 possessed a little bit wider band gap (2.71 eV) compared to 2.69 eV of g-C3N4. The VB maximum for pristine and P-doped g-C3N4 were found to be 2.07 eV, and 2.12 eV, respectively. The C/N molar ratio determined with EA was 0.79 and 0.85 for P-doped and pristine g-C3N4 that corresponded to non-perfect framework structures of both products. From the XPS analysis of C, N and P spectra the incorporation of P-atoms by the replacement of C-ones was concluded. XPS and 31P NMR data revealed the P atoms location at corner-carbon and bay-carbon sites of g-C3N4. The authors suggested the formation of electron-rich state of P-g-C3N4, which was confirmed by the measurement of zeta potentials of aqueous sample suspensions. After the P-doping, the charge of surface was changed from +7.3 mV (for g-C3N4) to −33.5 mV. The negatively charged surface of P-g-C3N4 promoted the adsorption of the cationic dye RhB. The authors suggested the formation of Lewis acid sites (P+ centers) and intrinsic Lewis base sites (amine or imine groups) (Figure 8) in g-C3N4 should promote the rapid separation of photogenerated electrons and holes and, thus, favoring the photocatalytic activity of H2 evolution and RhB degradation.
Table 5. Synthetic methods, applications, and photocatalytic efficiency of P-doped g-C3N4.
Table 5. Synthetic methods, applications, and photocatalytic efficiency of P-doped g-C3N4.
PrecursorSynthetic MethodC/Doping Element Atomic Ratio,
Doped (Pristine)
Photocatalytic ProcessConditions of the ProcessEfficiency Doped/PristineReferences
g-C3N4,
P powder
Thermal modificationP 2p signal at around 133.7 eVH2 evolution300 W Xe lamp (λ > 300 nm), filter λ > 420 nm, 10% vol% TEOA, Pt co-catalyst (3 wt.%)λ > 300 nm
261.2 µmol·g−1·h−1/81.6 µmol·g−1·h−1
λ > 420 nm
171.6 µmol·g−1·h−1/81.6 µmol·g−1·h−1
[109]
Urea, phosphonitrilic chloride Thermal condensation 4.4 atomic% by EDS
5.72 atomic% by XPS
H2O2 generationVisible light irradiation (420 nm ≤ λ ≤ 700 nm)1968 μmol·g−1·h−1/
68 μmol·g−1·h−1
[108]
NH4SCN, NH4PF6 Thermal condensation RhB destruction300 W Xe lamp, filter λ > 420 nm0.09856 min−1/
0.03679 min−1
[110]
g-C3N4,
phosphorene
Mechanically mixing1.8 wt.%H2 evolution300 W Xe lamp, filter λ > 400 nm, lactic acid (88 vol%)571 µmol·g−1·h−1/
43 μmol·g−1·h−1
[117]
g-C3N4, sodium hypophosphite Thermal treatment method 13.52 wt.%RhB destruction300 W Xe lamp,
λ: 420–780 nm,
0.0525 min−1/
0.0126min−1
[115]
Urea, adenosine phosphate Thermal condensation followed by thermal exfoliation method 2.17 atomic%H2 evolution300 W Xe lamp, filters
λ: 400, 420, 435, 450, 550 nm; 10% vol% TEOA, Pt co-catalyst (3 wt.%)
9523.7 µmol·g−1·h−1/
458 μmol·g−1·h−1
[118]
Urea, NH4H2PO2 Thermal condensation H2 evolution300 W Xe lamp, filter
λ 400 nm, 20% vol% TEOA, Pt co-catalyst (1 wt.%)
5.7 times that of pristine[119]
Dicyandiamide, NH4Cl, (NH4)2HPO4 Thermal condensation 1.53 wt% by EDSH2 evolution300 W Xe lamp, filter
λ 420 nm, 10% vol% TEOA, Pt co-catalyst (3 wt.%)
33.2 µmol·g−1·h−1/
10.7 μmol·g−1·h−1
[126]

7. Vacancy-Doped g-C3N4

The alternative way to modulate the g-C3N4 electronic band structure to favor visible light harvesting, accelerate charge separation, and transport was reported to be homogeneous self-modification with nitrogen [127,128,129,130,131,132,133,134,135,136,137,138] or carbon [88,139,140,141] vacancies. Native point defect doping via thermal treatment is an easy and promising method to tune the electrical transport properties of semiconductors made for renewable-energy conversion. By introducing additional energy levels and acting as reactive sites, vacancies can play a significant role in medication of the properties of photocatalysts (Table 6). The most common strategies of synthesis of nitrogen or carbon defective g-C3N4 are the post-heat treatment in H2 or Ar [127,141], the polymerization in various atmospheres [88,128,129], the post-heat treatment with a reducing agent [88,130,131], the thermal copolymerization with agents that deliberates gas [132,133,138], and the quick post-thermal treatment [135,136,139,142].
Wang et al. created N vacancies in g-C3N4 by the copolymerization of urea with oxalyl dihydrazide (ODH) as an environmental atmosphere control agent [133]. The thermal destruction of ODH caused the H2 liberation and produced a certain amount of heat that resulted in the defects formation in g-C3N4. The source of H2 was N2H4 as a product of the pyrolysis of ODH and NH3 (NH3 was produced by the thermal condensation of urea) at 135–140 °C. The SSA of pristine and modulated with N vacancies g-C3N4 (V-g-C3N4) were 85.4 and 116.9 m2·g−1, respectively. The progressive increasing of the C/N atomic ratio was observed. The atomic ratio of C/N determined with EA 0.65 for pristine and 0.74 for V-g-C3N4 was in agreement with the data from EDS and XPS (Table 4). The formation of C-C bond was confirmed using 13C solid-state magic-angle spinning nuclear magnetic resonance (MAS NMR) (Figure 9a) by the appearance of a small peak at 169.5 ppm in g-C3N4 modulated with N vacancies that was ascribed to C2N–C (3). The enhanced ESR signal at g = 1.9997 (Figure 9b) illustrated the presence of unpaired electrons on the C atoms, which appeared due to the loss of N atoms (Figure 9c).
Introduction of N vacancies resulted in increasing of the crystallinity of g-C3N4 as was detected by the XRD study. The authors also observed that the absorption band of the N vacancies doped g-C3N4 was red shifted compared with pristine g-C3N4, resulting in the bandgap narrowing from 2.75 eV to 2.61 eV. The PL, photocurrent, and EIS studies demonstrated that the interface charge transport can be effectively enhanced by the introduction of N vacancies.
The introduction of intra- and inter-triazine N-vacancies into g-C3N4 was performed under melamine treatment in various atmospheres by Li et al. [128]. The theoretical DFT calculation and experimental results allowed these authors to determine the type of N-vacancies obtained under various conditions. The inter-triazine N-vacancies were introduced after treating melamine in the CCl4 atmosphere. The polymerization of melamine in Ar and H2 resulted in the formation of “vacancy-free” and intra-triazine N-vacancies contained g-C3N4, respectively. The DFT calculation and optical properties study demonstrated that the bandgap was only slightly reduced by the intra-triazine N-vacancy but was strongly reduced by the inter-triazine N-vacancy. The band gap values of g-C3N4 obtained in CCl4, Ar, and H2 were determined to be 2.76, 2.79, and 2.80 eV, respectively. The minor changes in the values of SSA were observed for all samples: 11.2, 12.6, and 16.0 m2·g−1 for g-C3N4 treated at CCl4, Ar, and H2, respectively. Compared with “vacancy-free” g-C3N4, the EPR signals of g-C3N4 with inter- and intra-triazine N-vacancies were much enhanced, indicating the existence of more unpaired electrons. The efficient separation of photogenerated carriers was confirmed by the PL quenching and photocurrent enhancement of both types of nitrogen V-g-C3N4 due to the extra channel of electron transfer.
Xie et al. designed g-C3N4 with two types of regulatable nitrogen vacancies in a one pot method by the KOH-assisted calcination treatment of pristine g-C3N4 for a remarkably high H2O2 evolution photoactivity [131]. During the thermal treatment, the formation of N2C and NHx vacancies occurred due to changing molecular structure influenced by KOH. The loss of g-C3N4 ordered structure was caused by the treatment as well. The experimental analyses and theoretic DFT calculations revealed that the N2C vacancies improved the photoexcited charges separation and the NHx vacancy activated oxygen in two-electron process. The favorable amount of two types of N vacancies in g-C3N4 was determined. The N/C atomic ratios for a sample with an optimal amount of both types of N vacancies decreased compared with pristine g-C3N4. The SSA was 71 m2 g−1 and the thermal post treatment of g-C3N4 with and without KOH decreased from 48 to 45 m2 g−1, respectively. Due to the efficient carrier separation and oxygen activation processes, the nitrogen defective g-C3N4 demonstrated a 15-fold enhancement of H2O2 evolution (152.6 μmol·h−1) and superior stability over 52 h.
The carbon vacancies were prepared in the g-C3N4 framework via magnesium vapor etching by Li et al. [88]. They calcined a mixture of pristine g-C3N4 and magnesium powders to get magnesium vapor that etched the N-(C)3, C–N=C, or C=C lattice sites resulting in the vacancies. The applied method allowed the authors to increase the SSA from 58.5 m2 g−1 of g-C3N4 to 70.6 m2 g−1 of V-g-C3N4 that can supply more active centers for photocatalytic process. The decreased C/N atomic ratio from 0.78 of g-C3N4 to 0.51 of V-g-C3N4 was determined by the EDS analysis. The further evidence of carbon vacancies in g-C3N4 framework after etching was confirmed by the XPS analysis. The smaller C/N atomic ratio of V-g-C3N4 and the disappearance of the deconvoluted peak of C-C group in the high-resolution C 1s spectrum of V-g-C3N4 evidenced the origin of C vacancies from the destruction of C–C groups. A slight blue-shift of the absorption threshold and the absorption intensity enhancement from 450 nm to 800 nm was observed by DRS of V-g-C3N4. The calculated band-gap values were 2.87 eV and 2.98 eV for pristine and C vacancies-doped g-C3N4. The significant decrease of the V-g-C3N4 PL compared to pristine g-C3N4 confirmed its ability to inhibit charge recombination due to the capture of electrons by C vacancies. V-g-C3N4 also demonstrated the higher charge separation efficiency. The transient photocurrent responses under irradiation in an on-and-off cycle mode of carbon V-g-C3N4 was larger than that of pristine g-C3N4. As a result, V-g-C3N4 demonstrated the significant improvement of photocatalytic H2 generation performance.
Zhang et al. prepared porous ultrathin carbon defective g-C3N4 by the thermal treatment of g-C3N4 in two steps [139]. The obtained V-g-C3N4 possessed lower crystallinity compared to pristine g-C3N4 due to the formation of vacancies in the C3N4 structure. The two-step thermal treatment promoted the formation mesoporous structure with the increased SSA from 2.06 m2 g−1 to 162.57 m2 g−1. The formation of carbon defects was evidenced by the reduced C/N ratio from 0.85 to 0.81 determined by XPS. The enhanced EPR signal of V-g-C3N4 corresponded to the presence of C vacancies. The absorption edge of V-g-C3N4 was blue-shifted due to thermal oxidation of pristine g-C3N4. The band gap values were 2.74 and 3.02 eV for the pristine and carbon defective g-C3N4, respectively. The VB potentials of pristine and thermally exfoliated carbon V-g-C3N4 (determined by XPS) were 2.24 eV and 2.40 eV, respectively. The conduction bands were −0.50 eV and −0.62 eV, respectively. As the reduction potential of N2 to NH4+ was more positive than the CB potentials of both g-C3N4, the authors predicted the ability to reduce nitrogen under irradiation. The PL study of pristine g-C3N4 and V-g-C3N4 was not in agreement with other reports on doped g-C3N4 [88,139]. The authors attributed the longer lifetime of excited state of V-g-C3N4 to the decreasing of the recombination rate of electron–hole pairs. The confirmation of efficient charge carrier separation was obtained by the strong increasing of photocurrent response intensity of carbon V-g-C3N4 compared to g-C3N4. The capacity of V-g-C3N4 to transport charge carriers was suggested to be due to its porous structure as a result of carbon vacancies, which also promoted nitrogen fixation in a water environment.

8. Photocatalytic Applications of Non-Metal Elements Doped g-C3N4

Graphitic carbon nitride as a promising photocatalyst found applications in many photocatalytic processes, such as CO2 reduction, hydrogen evolution, pollutants degradation, and so forth [124,142,143,144,145,146,147]. The variety of strategies are offered to optimize photocatalytic processes by the modulation of g-C3N4-based photocatalysts composition and morphology to enlarge the range of absorbed light, promote charge carrier separation, and increase specific surface area.

8.1. Photocatalytic CO2 Reduction

The nanostructured graphitic carbon nitrides were suggested to be the perspective materials for CO2 capture and transformation into fine chemicals due to perfect semiconducting band gap with suitable band edges (Figure 10a–c).
The hierarchical O-g-C3N4 nanotubes obtained by thermal polycondensation of melamine exhibited significantly improved the photocatalytic CO2 reduction performance under visible light irradiation in comparison with pristine g-C3N4 [43]. Compared with pristine g-C3N4, O-g-C3N4 nanotubes have the following advantages: 1) Larger SSA and more photoactive sites; 2) wider visible-light absorption range and suitable band structure; 3) improved separation efficiency of charge carriers; and 4) high CO2 uptake capacity. The photocatalytic activity was evaluated by the generation rates of CH3OH as it was the main product of photocatalytic performance on both g-C3N4 and the nanotubes. The average CH3OH generation rate of 0.88 µmol·g−1·h−1 was about five times higher than that produced by g-C3N4 (0.17 µmol·g−1·h−1). The CH3OH generation rate was approximately the same in three cycles (3 h). Thus, it was concluded that the highly active and stable O-g-C3N4 nanotubes are the promising photocatalysts for hydrocarbon fuel generation by the photocatalytic CO2 reduction under visible light irradiation.
The direction for the design and synthesis of low-cost N-g-C3N4 materials for the photocatalytic CO2 reduction under UV light was proposed by Mo et al. [62]. As photocatalysts they used N-g-C3N4 nanotubes obtained by the heating of supramolecular intermediate under NH3, air, Ar, and N2 atmospheres. The influence of the porous structures and surface amino groups on the photocatalytic process was investigated. The main reaction products were determined to be CO and a small amount of H2. The highest CO formation rate of 103.6 μmol·g−1·h−1 was observed for pristine g-C3N4 (NH3) that was obtained by the precursor thermal treatment in the NH3 atmosphere. This rate was 17.0, 1.8, 2.7, and 2.8 times higher than that with g-C3N4 and the N-g-C3N4 materials, which were obtained by the thermal treatment in air, Ar, and N2, respectively. The high photocatalytic activity of g-C3N4 (NH3) catalyst was explained by the peculiar porous nanotube structure and by the increased quantity of amino groups. The porous 1D tubular structure improved the charge separation efficiency and provided numerous active sites for the surface reaction. The modification of g-C3N4 with amino groups not only extended the lifetime of excited species, but also caused the strong Lewis basicity, which was beneficial for the CO2 adsorption with further promotion of the CO2 photoreduction. The CO formation rates with g-C3N4 (NH3) photocatalyst was 15 and 13 times higher than those of TiO2 (P25) and black phosphorus, respectively. The nitrogen-doped sp2-carbon (graphitic)-rich electrodes demonstrated high selectivity for the CO2 reduction reaction products [148]. The authors evidenced that the host structure of nitrogen dopants was crucial for the catalytic activity observed.
The photocatalytic CO2 reduction of amine-functionalized g-C3N4 obtained by its treatment with monoethanolamine was evaluated under UV light [58]. The amine-modified g-C3N4 demonstrated the considerable enhancement in the CO2 conversion in comparison with pristine g-C3N4. The main products were found to be CH3OH and CH4. The pristine g-C3N4 was reported to have the photocatalytic CH3OH-production rate of 0.26 μmol·h−1·g−1 and only trace amount of CH4, while the optimal amine-functionalized sample provided the similar CH3OH production rate of 0.28 μmol·h−1·g−1 and, further, the CH4-production rate of 0.34 μmol·h−1·g−1. Similar to Mo et al. [62], the high activity of the amine-functionalized photocatalyst was attributed to the enhancement of CO2 adsorption and destabilization ability of g-C3N4 after the amine-functionalization treatment. The authors assumed that CO2 was adsorbed on the surface of amine-functionalized g-C3N4 through acid-base interactivity with amino groups and formed HCO3. As HCO3 was more active than linear CO2 it promoted the formation of CH4 [149]. It was shown that under the same conditions, CO2 molecules formed CH3OH [98]. Similar to Mo et al. [62], it was established that the exorbitant surface amine functionalization caused the decrease of photocatalytic activity due to reducing surface-active sites.
S-g-C3N4 fabricated by the thermal polycondensation of thiourea demonstrated enhanced activity in the photocatalytic reduction of CO2 into hydrocarbon fuels under UV–Vis light irradiation in comparison with pristine g-C3N4 [98]. The main product of photocatalytic CO2 reduction was CH3OH. The CH3OH yield with S-g-C3N4 was determined to be 1.12 μmol·g−1, whereas for pristine g-C3N4 it was 0.81μmol·g−1. The photogenerated electrons captured in these defects can promote the charge transfer and separation. The inhibition of the electron–hole recombination and prolongation of the lifetime of charge carriers occurs.
Overall, the O, N, S, and P-doped g-C3N4 demonstrated the enhanced activity in the CO2 photocatalytic reduction by improving the energy band structure that allowed more effective light trapping in the visible light region of spectra. The nonmetal element doping caused the formation of an impurity level in g-C3N4, which favored the charge transfer. Moreover, the surface groups, which are introduced by the doping process, enhanced the CO2 adsorption due to acid–base interactivity.

8.2. H2-Evolution

The graphitic carbon nitride presents a great potential for realizing hydrogen evolution reactions. The introduction of non-metal elements caused the exfoliation of bulk g-C3N4 into ultrathin nanosheets, activated the basal plane of g-C3N4, and improved the intrinsic electronic conductivity, thus implementing the facilitated H2 production.
The activity of the C-g-C3N4 composite fabricated by Cao et al. was evaluated by the production of H2 under visible light irradiation and compared with pristine g-C3N4 [71]. The H2 evolution rate using C-g-C3N4 (8.88 μmol/h) was 5.2 times higher that of g-C3N4 (1.70 μmol·h−1) due to the improvement of separation of photoexcited carriers and the enhanced utilization of visible light efficiency. The advantages of C-g-C3N4 were responsible for its drastic photocatalytic activity, such as the ability to separate more photoinduced electrons of CB and holes of VB under irradiation and the presence of delocalized π bonds, which reduce the barrier of charge carrier transfer and enhance electron transportation (Figure 11). The bridging nitrogen atoms of pristine g-C3N4 are unable to form conjugation with ambient carbon atoms due to inconsequent and small area of π conjugated system in pristine g-C3N4. Thus, under irradiation of C-g-C3N4, more and faster electrons were formed to participate in the process of hydrogen evolution.
O-g-C3N4 with the various content of oxygen was tested for the photocatalytic hydrogen evolution reactions to study the dependence of photocatalytic activity of O-g-C3N4 on oxygen coverage [24]. The reaction was performed at atmospheric pressure using N2 as a gas carrier in the presence of a hole scavenger and Pt cocatalyst under irradiation by a Xe lamp fitted with a filter to simulate solar light. The increasing of the O-g-C3N4 photocatalytic activity was observed in the sequence of surface O at.% compositions 3.6 > 3. 2> 0 > 5.4 > 6.6. The authors observed a competing dual effect of oxygen doping on the photocatalytic activity. On one hand, the enhanced light harvesting capability was influenced by the higher porosity of O-g-C3N4 and sub-gap impurity states formed by oxygen in its electronic band structure. Although, at the lower oxygen levels the dopant effectively captured and mediated the interfacial charge transfer that promoted the H2 evolution. On the other hand, the impurity levels from high-density oxygen groups become detrimental to the photoactivity of O-g-C3N4. These levels acted as recombination centers due to their positions closer to central position of the band gap.
O-g-C3N4 obtained by the thermal polymerization of a purposefully designed precursor with increased photoreduction capability demonstrated the notable enhancement of hydrogen evolution under visible light irradiation [42]. The average hydrogen evolution rate was 64.30 μmol·h−1, which is 17.8 times higher in comparison with 3.60 μmol/h of g-C3N4.
C60-g-C3N4 nanowire composites obtained by Long et al. through the calcination of urea and a C60 nanorods mixture demonstrated the higher activity in the photocatalytic H2 evolution process compared to pristine g-C3N4 [76]. Due to the band gap of g-C3N4 reduced by incorporation of C60, the charge carriers of C60-g-C3N4 nanowire composites became easily excited with visible light. The delocalized π structure of C60 promoted the photoinduced electron transfer and acted as effective electron acceptor. The irradiation of C60-g-C3N4 composites caused the CB photogenerated electrons transfer to C60, which resulted in the effective hole–electron separation. These separated charges were trapped by surface adsorbates to produce highly reactive radicals. The proton reduction to H2 occurred with electrons accumulated on C60, thus optimizing the H2 production. The H2 evolution rate for the C60-g-C3N4 nanowire composite was about 4.7 times higher than for pristine g-C3N4. The hydrogen was produced with the rate of 8.73 μmol/h at 5.10% of quantum efficiency in the presence of C60-g-C3N4.
The H2-evolution under visible light-inducing water splitting was performed to study the photocatalytic activity of N-g-C3N4 [63]. The authors used 3 wt.% Pt as a co-catalyst and 10 vol% triethanolamine as a hole sacrificial agent. The hydrogen evolution performance of N-g-C3N4 was 4.3 times higher in comparison with pristine g-C3N4. The doped N atoms were supposed to promote the separation and transfer of charge carriers and inhibit their recombination. N-g-C3N4 demonstrated its high stability after four cycles of the photocatalytic experiments.
Guo et al. suggested the improved optical absorption properties of N-g-C3N4 that was obtained by the copolymerization of urea with DMF and the decreased band-gap facilitated harvesting of visible light and increased photocatalytic activity [51]. The more negative conduction band potential of obtained N-g-C3N4 resulted in the improved reducing ability of photoelectrons, thus promoting the photocatalytic water splitting for H2 evolution. The rate of photocatalytic hydrogen production was 14 times higher than that of pristine g-C3N4.
Tian et al. studied the photocatalytic performance of mesoporous N-g-C3N4 materials obtained via the pre-hydrothermal treatment of urea and melamine for the photocatalytic H2 production under visible light irradiation [64]. The authors used 20 vol% lactic acid as a sacrificial agent and 1% Pt as a co-catalyst. The increased H2 evolution activity was observed for all the doped samples. The positive effect of melamine phase-transformation was observed even for N-g-C3N4 obtained without urea that demonstrated the photocatalytic H2 production rate of 1.7 times higher than that of pristine g-C3N4. The highest photocatalytic performance with the H2 evolution rate of 3579 μmol·h−1·g−1 was observed for the sample with the molar ratio of urea: melamine = 3:1, which was 23 times higher than that of g-C3N4. Consistent results were also obtained for the H2 evolution in the absence of Pt while maintaining other conditions to be similar. The sample with the molar ratio of urea: melamine = 0:1 and 3:1 demonstrated the H2 evolution rate of 1.9 and 15 times higher than that of pristine g-C3N4. The authors identified three factors that caused the high photocatalytic activity of the 3D ultrathin porous N-g-C3N4 materials: Firstly, the increased specific surface area originated from porous nanosheets that contained more reaction sites. Secondly, the narrowed band gap allowed more visible-light absorption and improved the H2 evolution ability. Thirdly, the significant reduction of recombination rate of photogenerated electrons and holes by porous structure of N-g-C3N4 improved the charge separation and rapid movement of charge carriers to the surface of photocatalyst where the reduction process occurred.
N vacancies-doped g-C3N4 with increased crystallinity and boosted visible-light harvesting demonstrated improved performance of the photocatalytic H2 evolution [133]. The hydrogen evolution rate 5833.1 μmol·h−1·g−1 compared to 1447.8 μmol·h−1·g−1 of pristine g-C3N4. The V-g-C3N4 photocatalyst had high chemical stability even after five cycles of the continuous photocatalytic reaction.
The theoretical and experimental study of the effect of intra- and inter-triazine N-vacancies in g-C3N4 on the photocatalytic H2 evolution activity was carried out by Li et al. [88]. The singly occupied defect states, which were formed in the band gap of g-C3N4 by both types of N vacancies, trapped the photogenerated electrons and served as centers of the H+ reduction. The most efficient H2 evolution was observed for V-g-C3N4 with the inter-triazine vacancies compared to V-g-C3N4 with the intra-triazine vacancies due to stronger electron localization. The normalized reaction rate of inter-triazine V-g-C3N4 was 9 times higher than that of “vacancy-free” g-C3N4, and 2.2 times higher than that of intra-triazine V-g-C3N4.
The SM-g-C3N4 nanosheets demonstrated enhanced activity in the photocatalytic H2 production compared to pristine g-C3N4 [103]. The average H2 evolution rate was 3.80 times higher than that of g-C3N4. The study of H2 production under irradiation with the wavelengths of 420, 450, 475, and 520 nm demonstrated that the activity of SM-g-C3N4 corresponded to the optical absorption of the photocatalyst due to the band gap excitation. The only trace amount of H2 was produced with pristine g-C3N4 when the irradiation light with the wavelength longer than 450 nm was applied.
In other research, the drastic increase of H2 evolution was reached with S-g-C3N4 nanosheets [90]. The obtained S-doped materials possessed the increased SSA, higher efficiency in charge carrier separation, and enlarged band gap energy with the upward shift of the CB potential and the downward shift of VB potential. These factors caused the high H2 production rate of 127.4 μmol·h−1, which is about 250 times higher than that for pristine g-C3N4. There was a stronger reduction ability observed for the hydrogen evolution of electrons in the CB of S-g-C3N4.
The great increase of photocatalytic activity in the H2 evolution under the visible-light irradiation was reached for the porous P-g-C3N4 nanosheets [118]. The total H2 production for 4 h was 1813 and 30,281 μmol for pristine g-C3N4 and P-g-C3N4, respectively. The photocatalytic H2 evolution rate was 458 and 9524 μmol−1·h−1, respectively. The obtained P-g-C3N4 was stable over the five test cycles. The apparent quantum efficiency was calculated to be 5.49%, 2.67%, 1.01%, and 0.17% at irradiation with λ = 400, 420, 450, and 550 nm, respectively. The enhancing of photocatalytic activity was reached due to the increased number of active sites for the photocatalytic H2 evolution provided by the porous structure and P-doping, the extending of light absorption range, and the improvement of hydrophilicity for easier water molecules adsorption via introducing OH groups. Additionally, the P-doping changed the excitation process and improved the efficiency of the charge separation.
The O, N, S, and P doping is a useful strategy to modify the electronic structure of g-C3N4 and to enhance the photocatalytic effect of hydrogen production. When the non-metal elements were applied to dope g-C3N4, the H2 generation was increased due to the lowering of the charge recombination rate and accelerated charge mobility.

8.3. Degradation of Dyes and Organic Pollutants

The photocatalytic activity of O-g-C3N4 was evaluated by the degradation of organic dyes Rhodamine B (RhB) and Methyl Orange (MO) under irradiation with simulated solar light [35]. Ninety-five percent of RhB (10 mg/L) was degraded in 20 min and 70% of MO (10 mg/L) in 4 h in the presence of O-g-C3N4. The determined rate constants were 65 times and 24 times higher than those of g-C3N4, for RhB and MO degradation, respectively. The significant improvement of photocatalytic activity was explained by the porous layered structure of O-g-C3N4, its ability to reduce the recombination of photogenerated carriers, and increased absorption of visible light. Among the possible active species (O2•−, OH, h+) O2•− and h+ were main active species of the photocatalytic degradation of RhB that was confirmed by using sacrificial agents.
The photocatalytic decomposition of RhB under visible light was performed to determine the influence of oxidative treatment of g-C3N4 with PMS [17]. The dependence of the photocatalytic activity of the resulting O-g-C3N4 on the amount of oxidant was found. Huang et al. determined the optimal amount of PMS for obtaining the O-g-C3N4 photocatalyst with the highest activity for the decomposition of RhB under visible light. The mechanism of RhB decomposition is not the same as for non-selective degradation of the chromophore in RhB by hydroxyl radicals described in the papers [34,35]. It was found that the RhB photocatalytic degradation by O-g-C3N4 occurred through N-deethylation pathways. The study of the RhB degradation in the presence of various inhibitors of photoactive species formation confirmed that O2•− was the dominant oxidant in the studied photocatalytic system (Figure 12).
A dual oxygen group (C-O-C and C=O) doped carbon nitride prepared by a two-step thermal treatment process demonstrated the remarkably enhanced visible-light photocatalytic performance for pollutant degradation [19]. Compounds, such as bisphenol, phenol, 2-chlorphenol, and diphenhydramine, were decomposed in their aqueous solutions using g-C3N4. The improvement of the degradation rate occurred due to the formation of an internal electric field and strong interactions between the photocatalyst and organic contaminant. The authors proposed a possible electron transfer pathway for enhanced pollutant photodegradation. Photoexcited electrons moved to the dual oxygen groups together with further generation of O2•− that decomposed the pollutants. The pollutant adsorption caused the enhancement of internal electric field and greatly increased photogenerated e/h+ separation and transfer. As a result, the oxidation with h+ and the generation of O2•− became more facile.
In other research, 49.3 times higher photocatalytic degradation efficiency of 4-nitrophenol using S-g-C3N4 than that with pristine g-C3N4 under visible light irradiation (>400 nm) was demonstrated [107]. The enhanced photocatalytic activity was attributed to the nanosheets formation and sulfur modification. Chrysanthemum-like nanosheet structures of S-g-C3N4 presented more active sites and improved its photochemical response. The sulfur modification induced the energy structure changes and lattice distortion. It was confirmed that the photocatalytic degradation of 4-nitrophenol by S-g-C3N4 was greatly influenced by oxidative species, such as singlet oxygen.
The high photocatalytic activities toward removing organic pollutants were demonstrated for mesoporous C-g-C3N4 ultrathin nanosheets synthesized through a facile one-step thermal condensation method with an agar-melamine gel precursor [79]. The more efficient ability of the nanosheets to remove RhB, phenol, Bisphenol A, and Phenanthrene compared to bulk g-C3N4 under visible-light irradiation was reported. The C-g-C3N4 ultrathin nanosheets demonstrated improved photoelectrochemical properties due to the strong enhancing of the photo-excited carriers’ separation and migration. It was found that O2•− as well as OH radicals were responsible for the strong oxidative capacity of the carbon doped photocatalysts, while for the pristine g-C3N4 only O2•−radicals oxidized the pollutants. The abundant holes of the VB h+ directly oxidized the pollutants. The great improvement photocatalytic activity was assigned to synergy of doped carbon and ultrathin nanosheets structure. The incorporated carbon strengthened visible-light absorbance, promoted charge separation, and prevented the photogenerated carriers’ recombination. The C-g-C3N4 nanosheets provided more active sites for photocatalytic reactions.
The photocatalytic performance of SM-g-C3N4 nanosheets under the visible light irradiation was tested for the degradation of dichlorotriazine dye used in textile crafts (Procion Red MX-5B) [103]. The highest activity in the process was observed for a sample with the highest SSA and narrowest bandgap. To determine the mechanism of the photocatalytic process the degradation of Procion Red MX-5B was performed with the most active photocatalysts in the presence of hole, OH, and O2•− scavengers. The most active species in the photocatalytic decomposition of the dichlorotriazine dye was determined to be the O2•−.
P-g-C3N4 obtained by the thermo-induced copolymerization method from a low cost and environmentally friendly P precursor were tested in the RhB degradation under irradiation with visible light. Due to significant changes in the electronic, surface, and semiconductor properties of g-C3N4 after P-doping the improved photocatalytic performance of RhB degradation was confirmed [126]. RhB in the solution (10 mg·L−1) was completely decomposed within 10 min while the decomposition of RhB solution of the same concentration lasted 30 min with pristine g-C3N4. The high ability of P-g-C3N4 to adsorb RhB also influenced the improvement of photocatalytic performance.
The presented research reports demonstrated that the photocatalytic activity of non-metal element-doped g-C3N4 has been greatly improved over pristine g-C3N4, and the degradation rates of organic dyes and pollutants have been enhanced to varying degrees. The doping of g-C3N4 with O, N, C, S, and P elements resulted in materials with stronger light absorbance, faster charge migration rate, and more photoreaction sites, providing the increased photocatalytic reaction rate. The comparison of the photocatalytic activity of doped g-C3N4 in the reported processes can be done generally by the comparison of the photocatalytic rate constants. Because absorption of photons plays a key role in photocatalytic processes, the apparent quantum yield must be used for reporting and comparing photocatalytic activities, which corresponds to lower limits of the quantum yield [150].

9. Conclusions

In this review, we summarized the recent research progress in the non-metal doping of graphitic carbon nitride used mostly for photocatalysis. The overview was focused on the modulation of composition and morphology of g-C3N4 with chemical doping with oxygen, sulfur, phosphor, nitrogen, carbon, as well as nitrogen and carbon vacancies. The efforts of various research groups to improve the photocatalytic activity of g-C3N4 in processes of the CO2 reduction, H2-evolution, and organic contaminants degradation by its non-metal doping were successful. Their results confirmed that g-C3N4 is a universal, environmentally friendly, low-price, and stable material for creating new photocatalysts with targeted properties.
One of targets of the incorporation of non-metallic dopants into g-C3N4 is the narrowing of the band gap, extending the sensitivity to the light of visible region as well as the reducing of recombination rate of electron–hole pairs for efficient charge separation that caused the enhancement of photocatalytic performances. However, in very few studies, a significant band gap reduction was observed after doping/modification with nonmetal atoms. In some cases, the efforts to modify pristine g-C3N4 caused even band gap enlargement. It is important to perform the doping of non-metal atoms into g-C3N4 matrices to reach the effective band gap narrowing. The surface doping can only cause the formation of some band gap localized states.
The prominent decreasing of the PL intensity of the majority of O-, C-, N-, S-, P-, and vacancy-doped g-C3N4 compared to the pristine g-C3N4 confirmed their ability to inhibit charge recombination due to the capture of electrons by the dopants. However, the study of PL decay is rarely used for g-C3N4 materials. Data on PL quenching can give quantitative characteristic of charge recombination inhibition processes. The creation of g-C3N4 materials with a strong visible light absorption and extremely low radiative recombination of excited charge carriers is still the target of future investigations.
It is well known that SSA has important characteristics of solid catalysts as it influences a number of reaction sites and the ability to absorb reaction products. The effect of active sites on the activity of g-C3N4-based photocatalysts is unquestionable. The higher SSA of the materials provides more active sites, and enhances the transfer and separation of charges in photocatalytic processes. The SSA of pristine g-C3N4 was strongly affected by synthesis methods. In some cases, the modification with non-metal atoms caused the increasing of SSA that enhanced their photocatalytic activity.
It should be mentioned that the relation between acid–base properties of g-C3N4 and non-metal elements doping and its influence on photocatalytic activity was not enlightened enough in the literature. The Bronsted and Lewis base centers are formed in the frame of g-C3N4 by uncondensed primary and tertiary amino groups, and aromatic amino groups of the three-s-triazine moieties. The presence of Lewis acid and base centers affected its application as active materials for dark catalysis and photocatalysis. The study of the connection between kinds of non-metal element doping and the acid–base properties of doped g-C3N4 will allow us to develop new materials with targeted properties for the appropriate photocatalytic process.

Author Contributions

H.S. wrote the manuscript, P.P. finalized the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Czech Science Foundation (project No. 19-15199S), EU structural funding in Operational Program Research, Development and Education (project No. CZ.02.1.01/0.0/0.0/16_019/0000853 „IET-ER“).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef] [PubMed]
  2. Dong, G.; Zhang, Y.; Pan, Q.; Qiu, J. A fantastic graphitic carbon nitride (g-C3N4) material: Electronic structure, photocatalytic and photoelectronic properties. J. Photochem. Photobiol. C Photochem. Rev. 2014, 20, 33–50. [Google Scholar] [CrossRef]
  3. Kroke, E. Novel group 14 nitrides. Coord. Chem. Rev. 2004, 248, 493–532. [Google Scholar] [CrossRef]
  4. 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]
  5. Hasija, V.; Raizada, P.; Sudhaik, A.; Sharma, K.; Kumar, A.; Singh, P.; Jonnalagadda, S.B.; Thakur, V.K. Recent advances in noble metal free doped graphitic carbon nitride based nanohybrids for photocatalysis of organic contaminants in water: A review. Appl. Mater. Today 2019, 15, 494–524. [Google Scholar] [CrossRef]
  6. Zhou, Z.; Zhang, Y.; Shen, Y.; Liu, S.; Zhang, Y. Molecular engineering of polymeric carbon nitride: Advancing applications from photocatalysis to biosensing and more. Chem. Soc. Rev. 2018, 47, 2298–2321. [Google Scholar] [CrossRef]
  7. Wang, A.; Wang, C.; Fu, L.; Wong-Ng, W.; Lan, Y. Recent Advances of Graphitic Carbon Nitride-Based Structures and Applications in Catalyst, Sensing, Imaging, and LEDs. Nano-Micro Lett. 2017, 9, 47. [Google Scholar] [CrossRef]
  8. Masih, D.; Ma, Y.; Rohani, S. Graphitic C3N4 based noble-metal-free photocatalyst systems: A review. Appl. Catal. B Environ. 2017, 206, 556–588. [Google Scholar] [CrossRef]
  9. Fu, J.; Yu, J.; Jiang, C.; Cheng, B. g-C3N4-Based Heterostructured Photocatalysts. Adv. Energy Mater. 2018, 8, 1701503, (1–31). [Google Scholar] [CrossRef]
  10. Zhao, Z.; Sun, Y.; Dong, F. Graphitic carbon nitride based nanocomposites: A review. Nanoscale 2015, 7, 15–37. [Google Scholar] [CrossRef]
  11. Chen, W.; Jiang, D.; Zhu, M.Y.; Shi, T.Y.; Li, H.N.; Wang, K. An effective strategy for fabricating highly dispersed nanoparticles on O-C3N4 with enhanced electrocatalytic activity and stability. J. Alloy. Compd. 2018, 741, 1203–1211. [Google Scholar] [CrossRef]
  12. Yousefi, M.; Villar-Rodil, S.; Paredes, J.I.; Moshfegh, A.Z. Oxidized graphitic carbon nitride nanosheets as an effective adsorbent for organic dyes and tetracycline for water remediation. J. Alloy. Compd. 2019, 809, 11. [Google Scholar] [CrossRef]
  13. Zhu, W.R.; Hao, N.; Lu, J.W.; Dai, Z.; Qian, J.; Yang, X.D.; Wang, K. Highly active metal-free peroxidase mimics based on oxygen-doped carbon nitride by promoting electron transfer capacity. Chem. Commun. 2020, 56, 1409–1412. [Google Scholar] [CrossRef] [PubMed]
  14. Chubenko, E.B.; Baglov, A.V.; Leonenya, M.S.; Yablonskii, G.P.; Borisenko, V.E. Structure of Photoluminescence Spectra of Oxygen-Doped Graphitic Carbon Nitride. J. Appl. Spectrosc. 2020, 87, 9–14. [Google Scholar] [CrossRef]
  15. Denisov, N.M.; Chubenko, E.B.; Bondarenko, V.P.; Borisenko, V.E. Synthesis of Oxygen-Doped Graphitic Carbon Nitride from Thiourea. Tech. Phys. Lett. 2019, 45, 108–110. [Google Scholar] [CrossRef]
  16. Gao, Y.W.; Zhu, Y.; Lyu, L.; Zeng, Q.Y.; Xing, X.C.; Hu, C. Electronic Structure Modulation of Graphitic Carbon Nitride by Oxygen Doping for Enhanced Catalytic Degradation of Organic Pollutants through Peroxymonosulfate Activation. Environ. Sci. Technol. 2018, 52, 14371–14380. [Google Scholar] [CrossRef]
  17. Huang, J.; Nie, G.; Ding, Y.B. Metal-Free Enhanced Photocatalytic Activation of Dioxygen by g-C3N4 Doped with Abundant Oxygen-Containing Functional Groups for Selective N-Deethylation of Rhodamine B. Catalysts 2020, 10, 6. [Google Scholar] [CrossRef] [Green Version]
  18. Huang, T.; Pan, S.G.; Shi, L.L.; Yu, A.P.; Wang, X.; Fu, Y.S. Hollow porous prismatic graphitic carbon nitride with nitrogen vacancies and oxygen doping: A high-performance visible light-driven catalyst for nitrogen fixation. Nanoscale 2020, 12, 1833–1841. [Google Scholar] [CrossRef]
  19. Li, F.; Han, M.E.; Jin, Y.; Zhang, L.L.; Li, T.; Gao, Y.W.; Hu, C. Internal electric field construction on dual oxygen group-doped carbon nitride for enhanced photodegradation of pollutants under visible light irradiation. Appl. Catal. B Environ. 2019, 256, 10. [Google Scholar] [CrossRef]
  20. Liu, C.Y.; Huang, H.W.; Cui, W.; Dong, F.; Zhang, Y.H. Band structure engineering and efficient charge transport in oxygen substituted g-C3N4 for superior photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2018, 230, 115–124. [Google Scholar] [CrossRef]
  21. Li, Q.; Wang, S.C.; Sun, Z.X.; Tang, Q.J.; Liu, Y.Q.; Wang, L.Z.; Wang, H.Q.; Wu, Z.B. Enhanced CH4 selectivity in CO2 photocatalytic reduction over carbon quantum dots decorated and oxygen doping g-C3N4. Nano Res. 2019, 12, 2749–2759. [Google Scholar] [CrossRef]
  22. Wei, F.Y.; Liu, Y.; Zhao, H.; Ren, X.N.; Liu, J.; Hasan, T.; Chen, L.H.; Li, Y.; Su, B.L. Oxygen self-doped g-C3N4 with tunable electronic band structure for unprecedentedly enhanced photocatalytic performance. Nanoscale 2018, 10, 4515–4522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Chen, H.; Yao, J.H.; Qiu, P.X.; Xu, C.M.; Jiang, F.; Wang, X. Facile surfactant assistant synthesis of porous oxygen-doped graphitic carbon nitride nanosheets with enhanced visible light photocatalytic activity. Mater. Res. Bull. 2017, 91, 42–48. [Google Scholar] [CrossRef]
  24. 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 gC(3)N(4) and its effects on photocatalytic activity. Appl. Surf. Sci. 2020, 504. [Google Scholar] [CrossRef]
  25. Sun, S.D.; Li, J.; Cui, J.; Gou, X.F.; Yang, Q.; Liang, S.H.; Yang, Z.M.; Zhang, J.M. Constructing oxygen-doped g-C3N4 nanosheets with an enlarged conductive band edge for enhanced visible-light-driven hydrogen evolution. Inorg. Chem. Front. 2018, 5, 1721–1727. [Google Scholar] [CrossRef]
  26. Chen, Y.L.; Liu, X.Q.; Hou, L.; Guo, X.R.; Fu, R.W.; Sun, J.M. Construction of covalent bonding oxygen-doped carbon nitride/graphitic carbon nitride Z-scheme heterojunction for enhanced visible-light-driven H2 evolution. Chem. Eng. J. 2020, 383. [Google Scholar] [CrossRef]
  27. Sun, Z.Z.; Wang, W.; Chen, Q.W.; Pu, Y.Y.; He, H.; Zhuang, W.M.; He, J.Q.; Huang, L.M. A hierarchical carbon nitride tube with oxygen doping and carbon defects promotes solar-to-hydrogen conversion. J. Mater. Chem. A 2020, 8, 3160–3167. [Google Scholar] [CrossRef]
  28. Yuan, X.J.; Xie, R.L.; Zhang, Q.; Sun, L.; Long, X.J.; Xia, D.S. Oxygen functionalized graphitic carbon nitride as an efficient metal-free ozonation catalyst for atrazine removal: Performance and mechanism. Sep. Purif. Technol. 2019, 211, 823–831. [Google Scholar] [CrossRef]
  29. Zeng, Y.X.; Liu, X.; Liu, C.B.; Wang, L.L.; Xia, Y.C.; Zhang, S.Q.; Luo, S.L.; 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]
  30. Zhang, C.; Zhang, M.Y.; Li, Y.; Shuai, D.M. Visible-light-driven photocatalytic disinfection of human adenovirus by a novel heterostructure of oxygen-doped graphitic carbon nitride and hydrothermal carbonation carbon. Appl. Catal. B Environ. 2019, 248, 11–21. [Google Scholar] [CrossRef]
  31. Wang, Y.X.; Wang, H.; Chen, F.Y.; Cao, F.; Zhao, X.H.; Meng, S.G.; Cui, Y.J. Facile synthesis of oxygen doped carbon nitride hollow microsphere for photocatalysis. Appl. Catal. B Environ. 2017, 206, 417–425. [Google Scholar] [CrossRef]
  32. Mei, R.G.; Ma, L.; An, L.A.; Wang, F.; Xi, J.J.; Sun, H.Y.; Luo, Z.K.; Wu, Q.X. Layered Spongy-like O-Doped g-C3N4: An Efficient Non-Metal Oxygen Reduction Catalyst for Alkaline Fuel Cells. J. Electrochem. Soc. 2017, 164, F354–F363. [Google Scholar] [CrossRef]
  33. Wang, C.; Fan, H.Q.; Ren, X.H.; Ma, J.W.; Fang, J.W.; Wang, W.J. Hydrothermally Induced Oxygen Doping of Graphitic Carbon Nitride with a Highly Ordered Architecture and Enhanced Photocatalytic Activity. Chemsuschem 2018, 11, 700–708. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, H.; Guan, Y.H.; Hu, S.Z.; Pei, Y.B.; Ma, W.T.; Fan, Z.P. 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. [Google Scholar] [CrossRef]
  35. Zhang, B.; Li, X.J.; Zhao, Y.; Song, H.; Wang, H. Facile synthesis of oxygen doped mesoporous graphitic carbon nitride with high photocatalytic degradation efficiency under simulated solar irradiation. In Colloids and Surfaces a-Physicochemical and Engineering Aspects; Elsevier: Amsterdam, The Netherlands, 2019; Volume 580. [Google Scholar] [CrossRef]
  36. Zhang, J.W.; Gong, S.; Mahmood, N.; Pan, L.; Zhang, X.W.; Zou, J.J. Oxygen-doped nanoporous carbon nitride via water-based homogeneous supramolecular assembly for photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2018, 221, 9–16. [Google Scholar] [CrossRef]
  37. Zhu, K.; Ouyang, J.; Liu, J.M.; Zhu, Y.X.; Zeng, Q.; Cui, Y.J. Preparation and Photocatalytic Hydrogen Evolution from Water of Oxygen Doped Carbon Nitride Nanosheets. Chin. J. Inorg. Chem. 2019, 35, 1005–1012. [Google Scholar] [CrossRef]
  38. Zhu, Y.; Chen, Z.H.; Gao, Y.W.; Hu, C. General synthesis of carbon and oxygen dual-doped graphitic carbon nitride via copolymerization for non-photochemical oxidation of organic pollutant. J. Hazard. Mater. 2020, 394. [Google Scholar] [CrossRef]
  39. Tang, R.; Ding, R.L.; Xie, X.C. Preparation of oxygen-doped graphitic carbon nitride and its visible-light photocatalytic performance on bisphenol A degradation. Water Sci. Technol. 2018, 78, 1023–1033. [Google Scholar] [CrossRef] [Green Version]
  40. Wang, Y.X.; Rao, L.; Wang, P.F.; Guo, Y.; Guo, X.; Zhang, L.X. Porous oxygen-doped carbon nitride: Supramolecular preassembly technology and photocatalytic degradation of organic pollutants under low-intensity light irradiation. Environ. Sci. Pollut. Res. 2019, 26, 15710–15723. [Google Scholar] [CrossRef]
  41. Zhang, S.; Liu, Y.; Gu, P.C.; Ma, R.; Wen, T.; Zhao, G.X.; Li, L.; Ai, Y.J.; Hu, C.; Wang, X.K. Enhanced photodegradation of toxic organic pollutants using dual-oxygen-doped porous g-C3N4: Mechanism exploration from both experimental and DFT studies. Appl. Catal. B Environ. 2019, 248, 1–10. [Google Scholar] [CrossRef]
  42. Song, P.; Liang, S.H.; Cui, J.; Ren, D.; Duan, R.Y.; Yang, Q.; Sun, S.D. Purposefully designing novel hydroxylated and carbonylated melamine towards the synthesis of targeted porous oxygen-doped g-C3N4 nanosheets for highly enhanced photocatalytic hydrogen production. Catal. Sci. Technol. 2019, 9, 5150–5159. [Google Scholar] [CrossRef]
  43. Fu, J.W.; Zhu, B.C.; Jiang, C.J.; Cheng, B.; You, W.; Yu, J.G. Hierarchical Porous O-Doped g-C3N4 with Enhanced Photocatalytic CO2 Reduction Activity. Small 2017, 13, 9. [Google Scholar] [CrossRef] [PubMed]
  44. Miller, D.R.; Wang, J.J.; Gillan, E.G. Rapid, facile synthesis of nitrogen-rich carbon nitride powders. J. Mater. Chem. 2002, 12, 2463–2469. [Google Scholar] [CrossRef]
  45. Miller, D.R.; Swenson, D.C.; Gillan, E.G. Synthesis and structure of 2,5,8-triazido-s-heptazine: An energetic and luminescent precursor to nitrogen-rich carbon nitrides. J. Am. Chem. Soc. 2004, 126, 5372–5373. [Google Scholar] [CrossRef]
  46. Miller, D.R.; Holst, J.R.; Gillan, E.G. Nitrogen-rich carbon nitride network materials via the thermal decomposition of 2,5,8-triazido-s-heptazine. Inorg. Chem. 2007, 46, 2767–2774. [Google Scholar] [CrossRef]
  47. Huynh, M.H.V.; Hiskey, M.A.; Archuleta, J.G.; Roemer, E.L.; Gilardi, R. 3,6-di(azido)-1,2,4,5-tetrazine: A precursor for the preparation of carbon nanospheres and nitrogen-rich carbon nitrides. Angew. Chem. Int. Edit. 2004, 43, 5658–5661. [Google Scholar] [CrossRef]
  48. Gillan, E.G. Synthesis of nitrogen-rich carbon nitride networks from an energetic molecular azide precursor. Chem. Mat. 2000, 12, 3906–3912. [Google Scholar] [CrossRef]
  49. Fang, J.W.; Fan, H.Q.; Li, M.M.; Long, C.B. Nitrogen self-doped graphitic carbon nitride as efficient visible light photocatalyst for hydrogen evolution. J. Mater. Chem. A 2015, 3, 13819–13826. [Google Scholar] [CrossRef]
  50. Xu, Q.X.; Xu, G.Q.; Yu, Q.B.; Yang, K.; Li, H.Q. Nitrogen self-doped high specific surface area graphite carbon nitride for photocatalytic degradating of methylene blue. J. Nanopart. Res. 2019, 21, 13. [Google Scholar] [CrossRef]
  51. Guo, F.; Wang, L.J.; Sun, H.R.; Li, M.Y.; Shi, W.L. High-efficiency photocatalytic water splitting by a N-doped porous g-C3N4 nanosheet polymer photocatalyst derived from urea and N,N-dimethylformamide. Inorg. Chem. Front. 2020, 7, 1770–1779. [Google Scholar] [CrossRef]
  52. Jiang, L.B.; Yuan, X.Z.; Zeng, G.M.; Liang, J.; Wu, Z.B.; Yu, H.B.; Mo, D.; Wang, H.; Xiao, Z.H.; Zhou, C.Y. 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]
  53. Wang, H.H.; Li, Q.; Zhang, S.; Chen, Z.S.; Wang, W.X.; Zhao, G.X.; Zhuang, L.; Hu, B.W.; Wang, X.K. Visible-light-driven N-2-g-C3N4 as a highly stable and efficient photocatalyst for bisphenol A and Cr(VI) removal in binary systems. Catal. Today 2019, 335, 110–116. [Google Scholar] [CrossRef]
  54. Dong, S.S.; Liu, C.; Chen, Y.G. Boosting exciton dissociation and molecular oxygen activation by in-plane grafting nitrogen-doped carbon nanosheets to graphitic carbon nitride for enhanced photocatalytic performance. J. Colloid Interface Sci. 2019, 553, 59–70. [Google Scholar] [CrossRef] [PubMed]
  55. Qi, H.L.; Liu, Y.N.; Li, C.Y.; Zou, X.H.; Huang, Y.D.; Wang, Y.G. Precursor-reforming protocol to synthesis of porous N-doped g-C3N4 for highly improved photocatalytic water treatments. Mater. Lett. 2020, 264, 4. [Google Scholar] [CrossRef]
  56. Che, H.N.; Che, G.B.; Zhou, P.J.; Liu, C.B.; Dong, H.J.; Li, C.X.; Song, N.; Li, C.M. Nitrogen doped carbon ribbons modified g-C3N4 for markedly enhanced photocatalytic H2-production in visible to near-infrared region. Chem. Eng. J. 2020, 382, 9. [Google Scholar] [CrossRef]
  57. Hao, Q.G.; Song, Y.H.; Ji, H.Y.; Mo, Z.; She, X.J.; Deng, J.J.; Muhmood, T.; Wu, X.Y.; Yuan, S.Q.; Xu, H.; et al. Surface N modified 2D g-C3N4 nanosheets derived from DMF for photocatalytic H2 evolution. Appl. Surf. Sci. 2018, 459, 845–852. [Google Scholar] [CrossRef]
  58. Huang, Q.; Yu, J.G.; Cao, S.W.; Cui, C.; Cheng, B. Efficient photocatalytic reduction of CO2 by amine-functionalized g-C3N4. Appl. Surf. Sci. 2015, 358, 350–355. [Google Scholar] [CrossRef]
  59. Yan, J.; Zhou, C.J.; Li, P.R.; Chen, B.H.; Zhang, S.S.; Dong, X.P.; Xi, F.N.; Liu, J.Y. Nitrogen-rich graphitic carbon nitride: Controllable nanosheet-like morphology, enhanced visible light absorption and superior photocatalytic performance. Colloids Surf. A Physicochem. Eng. Asp. 2016, 508, 257–264. [Google Scholar] [CrossRef]
  60. Gao, B.R.; Wang, J.; Dou, M.M.; Huang, X.; Yu, X.X. Novel nitrogen-rich g-C3N4 with adjustable energy band by introducing triazole ring for cefotaxime removal. Sep. Purif. Technol. 2020, 241, 9. [Google Scholar] [CrossRef]
  61. Wu, X.H.; Gao, D.D.; Wang, P.; Yu, H.G.; Yu, J.G. NH4Cl-induced low-temperature formation of nitrogen-rich g-C3N4 nanosheets with improved photocatalytic hydrogen evolution. Carbon 2019, 153, 757–766. [Google Scholar] [CrossRef]
  62. Mo, Z.; Zhu, X.W.; Jiang, Z.F.; Song, Y.H.; Liu, D.B.; Li, H.P.; Yang, X.F.; She, Y.B.; Lei, Y.C.; Yuan, S.Q.; et al. Porous nitrogen-rich g-C3N4 nanotubes for efficient photocatalytic CO2 reduction. Appl. Catal. B Environ. 2019, 256, 7. [Google Scholar] [CrossRef]
  63. Zhou, Y.J.; Zhang, L.X.; Huang, W.M.; Kong, Q.L.; Fan, X.Q.; Wang, M.; Shi, J.L. N-doped graphitic carbon-incorporated g-C3N4 for remarkably enhanced photocatalytic H2 evolution under visible light. Carbon 2016, 99, 111–117. [Google Scholar] [CrossRef]
  64. Tian, N.; Zhang, Y.H.; Li, X.W.; Xiao, K.; Du, X.; Dong, F.; Waterhouse, G.I.N.; Zhang, T.R.; Huang, H.W. Precursor-reforming protocol to 3D mesoporous g-C3N4 established by ultrathin self-doped nanosheets for superior hydrogen evolution. Nano Energy 2017, 38, 72–81. [Google Scholar] [CrossRef]
  65. Wu, M.; Gong, Y.S.; Nie, T.; Zhang, J.; Wang, R.; Wang, H.W.; He, B.B. Template-free synthesis of nanocage-like g-C3N4 with high surface area and nitrogen defects for enhanced photocatalytic H2 activity. J. Mater. Chem. A 2019, 7, 5324–5332. [Google Scholar] [CrossRef]
  66. Wang, F.L.; Chen, P.; Feng, Y.P.; Xie, Z.J.; Liu, Y.; Su, Y.H.; Zhang, Q.X.; Wang, Y.F.; Yao, K.; Lv, W.Y.; et al. Facile synthesis of N-doped carbon dots/g-C3N4 photocatalyst with enhanced visible-light photocatalytic activity for the degradation of indomethacin. Appl. Catal. B Environ. 2017, 207, 103–113. [Google Scholar] [CrossRef]
  67. Asadzadeh-Khaneghah, S.; Habibi-Yangjeh, A.; Nakata, K. Decoration of carbon dots over hydrogen peroxide treated graphitic carbon nitride: Exceptional photocatalytic performance in removal of different contaminants under visible light. J. Photochem. Photobiol. A Chem. 2019, 374, 161–172. [Google Scholar] [CrossRef]
  68. Deng, P.H.; Li, H.Y.; Wang, Z.D.; Hou, Y. Enhanced photocatalytic hydrogen evolution by carbon-doped carbon nitride synthesized via the assistance of cellulose. Appl. Surf. Sci. 2020, 504. [Google Scholar] [CrossRef]
  69. Duan, Y.; Deng, L.; Shi, Z.; Liu, X.; Zeng, H.X.; Zhang, H.J.; Crittenden, J. Efficient sulfadiazine degradation via in-situ epitaxial grow of Graphitic Carbon Nitride (g-C3N4) on carbon dots heterostructures under visible light irradiation: Synthesis, mechanisms and toxicity evaluation. J. Colloid Interface Sci. 2020, 561, 696–707. [Google Scholar] [CrossRef]
  70. Jeong, T.; Piao, H.; Park, S.; Yang, J.H.; Choi, G.; Wu, Q.; Kang, H.; Woo, H.J.; Jung, S.J.; Kim, H.; et al. Atomic and electronic structures of graphene-decorated graphitic carbon nitride (g-C3N4) as a metal-free photocatalyst under visible-light. Appl. Catal. B Environ. 2019, 256, 7. [Google Scholar] [CrossRef]
  71. Cao, J.S.; Fan, H.Q.; Wang, C.; Ma, J.W.; Dong, G.Z.; Zhang, M.C. Facile synthesis of carbon self-doped g-C3N4 for enhanced photocatalytic hydrogen evolution. Ceram. Int. 2020, 46, 7888–7895. [Google Scholar] [CrossRef]
  72. Chatzoudis, A.; Giannopoulos, V.; Hollmann, F.; Smonou, I. Surface-Doped Graphitic Carbon Nitride Catalyzed Photooxidation of Olefins and Dienes: Chemical Evidence for Electron Transfer and Singlet Oxygen Mechanisms. Catalysts 2019, 9, 639. [Google Scholar] [CrossRef] [Green Version]
  73. Ge, F.Y.; Xu, Y.G.; Zhou, Y.H.; Tian, D.; Huang, S.Q.; Xie, M.; Xu, H.; Li, H.M. Surface amorphous carbon doping of carbon nitride for efficient acceleration of electron transfer to boost photocatalytic activities. Appl. Surf. Sci. 2020, 507. [Google Scholar] [CrossRef]
  74. Li, Z.H.; Huang, D.L.; Zhou, C.Y.; Xue, W.J.; Lei, L.; Deng, R.; Yang, Y.; Chen, S.; Wang, W.J.; Wang, Z.W. Metal-free carbon nitride with boosting photo-redox ability realized by the controlled carbon dopants. Chem. Eng. J. 2020, 382. [Google Scholar] [CrossRef]
  75. Liu, H.P.; Liang, J.; Fu, S.; Li, L.; Cui, J.H.; Gao, P.H.; Zhao, F.Y.; Zhou, J.G. N doped carbon quantum dots modified defect-rich g-C3N4 for enhanced photocatalytic combined pollutions degradation and hydrogen evolution. Colloids Surf. A Physicochem. Eng. Asp. 2020, 591, 13. [Google Scholar] [CrossRef]
  76. Long, D.; Chen, W.L.; Rao, X.; Zheng, S.H.; Zhang, Y.P. Synergetic effect of C-60/g-C3N4 nanowire composites for enhanced photocatalytic H2 evolution under visible light irradiation. ChemCatChem 2020, 12, 2022–2031. [Google Scholar] [CrossRef]
  77. Meng, L.R.; Yin, W.H.; Wang, S.S.; Wu, X.G.; Hou, J.H.; Yin, W.Q.; Feng, K.; Ok, Y.S.; Wang, X.Z. Photocatalytic behavior of biochar-modified carbon nitride with enriched visible-light reactivity. Chemosphere 2020, 239. [Google Scholar] [CrossRef]
  78. Wang, X.F.; Cheng, J.J.; Yu, H.G.; Yu, J.G. A facile hydrothermal synthesis of carbon dots modified g-C3N4 for enhanced photocatalytic H2-evolution performance. Dalton Trans. 2017, 46, 6417–6424. [Google Scholar] [CrossRef] [PubMed]
  79. Wang, Y.M.; Cai, H.Y.; Qian, F.F.; Li, Y.M.; Yu, J.Q.; Yang, X.L.; Bao, M.T.; Li, X.M. Facile one-step synthesis of onion-like carbon modified ultrathin g-C3N4 2D nanosheets with enhanced visible-light photocatalytic performance. J. Colloid Interface Sci. 2019, 533, 47–58. [Google Scholar] [CrossRef] [PubMed]
  80. Huang, Y.Y.; Li, D.; Fang, Z.Y.; Chen, R.J.; Luo, B.F.; Shi, W.D. Controlling carbon self-doping site of g-C3N4 for highly enhanced visible-light-driven hydrogen evolution. Appl. Catal. B Environ. 2019, 254, 128–134. [Google Scholar] [CrossRef]
  81. Li, X.L.; Bai, J.F.; Li, J.Q.; Li, C.; Zhong, X.Y.; Deng, S.P. The effect of n-pi* electronic transitions on the N-2 photofixation ability of carbon self-doped honeycomb-like g-C3N4 prepared via microwave treatment. RSC Adv. 2020, 10, 7019–7025. [Google Scholar] [CrossRef] [Green Version]
  82. Ran, M.X.; Li, J.R.; Cui, W.; Li, Y.H.; Li, P.D.; Dong, F. Efficient and stable photocatalytic NO removal on C self-doped g-C3N4: Electronic structure and reaction mechanism. Catal. Sci. Technol. 2018, 8, 3387–3394. [Google Scholar] [CrossRef]
  83. Xu, Q.L.; Jiang, C.J.; Cheng, B.; Yu, J.G. Enhanced visible-light photocatalytic H2-generation activity of carbon/g-C3N4 nanocomposites prepared by two-step thermal treatment. Dalton Trans. 2017, 46, 10611–10619. [Google Scholar] [CrossRef] [PubMed]
  84. Xu, J.; Huang, J.; Wang, Z.P.; Zhu, Y.F. Enhanced visible-light photocatalytic degradation and disinfection performance of oxidized nanoporous g-C3N4 via decoration with graphene oxide quantum dots. Chin. J. Catal. 2020, 41, 474–484. [Google Scholar] [CrossRef]
  85. Bao, N.; Hu, X.D.; Zhang, Q.Z.; Miao, X.H.; Jie, X.Y.; Zhou, S. Synthesis of porous carbon-doped g-C3N4 nanosheets with enhanced visible-light photocatalytic activity. Appl. Surf. Sci. 2017, 403, 682–690. [Google Scholar] [CrossRef]
  86. Chen, Z.; Fan, T.T.; Yu, X.; Wu, Q.L.; Zhu, Q.H.; Zhang, L.Z.; Li, J.H.; Fang, W.P.; Yi, X.D. Gradual carbon doping of graphitic carbon nitride towards metal-free visible light photocatalytic hydrogen evolution. J. Mater. Chem. A 2018, 6, 15310–15319. [Google Scholar] [CrossRef]
  87. Luo, L.; Ma, J.N.; Zhu, H.X.; Tang, J.W. Embedded carbon in a carbon nitride hollow sphere for enhanced charge separation and photocatalytic water splitting. Nanoscale 2020, 12, 7339–7346. [Google Scholar] [CrossRef] [Green Version]
  88. Li, J.X.; Ma, W.Q.; Chen, J.J.; An, N.; Zhao, Y.; Wang, D.J.; Mao, Z.Y. Carbon vacancies improved photocatalytic hydrogen generation of g-C3N4 photocatalyst via magnesium vapor etching. Int. J. Hydrogen Energy 2020, 45, 13939–13946. [Google Scholar] [CrossRef]
  89. Ma, L.; Hu, S.Z.; Li, P.; Wang, Q.; Ma, H.F.; Li, W. In situ synthesis of sulfur doped carbon nitride with enhanced photocatalytic performance using DBD plasma treatment under H2S atmosphere. J. Phys. Chem. Solids 2018, 118, 166–171. [Google Scholar] [CrossRef]
  90. Lv, H.Q.; Huang, Y.; Koodali, R.T.; Liu, G.M.; Zeng, Y.B.; Meng, Q.G.; Yuan, M.Z. Synthesis of Sulfur-Doped 2D Graphitic Carbon Nitride Nanosheets for Efficient Photocatalytic Degradation of Phenol and Hydrogen Evolution. ACS Appl. Mater. Interfaces 2020, 12, 12656–12667. [Google Scholar] [CrossRef]
  91. Wang, W.J.; Zeng, Z.T.; Zeng, G.M.; Zhang, C.; Xiao, R.; Zhou, C.Y.; Xiong, W.P.; Yang, Y.; Lei, L.; Liu, Y.; et al. Sulfur doped carbon quantum dots loaded hollow tubular g-C3N4 as novel photocatalyst for destruction of Escherichia coli and tetracycline degradation under visible light. Chem. Eng. J. 2019, 378, 11. [Google Scholar] [CrossRef]
  92. Joseph, S.; Abraham, S.; Priyanka, R.N.; Abraham, T.; Suresh, A.; Mathew, B. In situ S-doped ultrathin gC(3)N(4) nanosheets coupled with mixed-dimensional (3D/1D) nanostructures of silver vanadates for enhanced photocatalytic degradation of organic pollutants. New J. Chem. 2019, 43, 10618–10630. [Google Scholar] [CrossRef]
  93. Sakthivel, A.; Chandrasekaran, A.; Jayakumar, S.; Manickam, P.; Alwarappan, S. Sulphur Doped Graphitic Carbon Nitride as an Efficient Electrochemical Platform for the Detection of Acetaminophen. J. Electrochem. Soc. 2019, 166, B1461–B1469. [Google Scholar] [CrossRef]
  94. Fang, Y.X.; Li, X.C.; Wang, Y.; Giordano, C.; Wang, X.C. Gradient sulfur doping along polymeric carbon nitride films as visible light photoanodes for the enhanced water oxidation. Appl. Catal. B Environ. 2020, 268, 6. [Google Scholar] [CrossRef]
  95. Fan, Q.J.; Liu, J.J.; Yu, Y.C.; Zuo, S.L.; Li, B.S. A simple fabrication for sulfur doped graphitic carbon nitride porous rods with excellent photocatalytic activity degrading RhB dye. Appl. Surf. Sci. 2017, 391, 360–368. [Google Scholar] [CrossRef]
  96. Lin, Y.R.; Dizon, G.V.C.; Yamada, K.; Liu, C.Y.; Venault, A.; Lin, H.Y.; Yoshida, M.; Hu, C.C. Sulfur-doped g-C3N4 nanosheets for photocatalysis: Z-scheme water splitting and decreased biofouling. J. Colloid Interface Sci. 2020, 567, 202–212. [Google Scholar] [CrossRef]
  97. Lu, C.H.; Zhang, P.; Jiang, S.J.; Wu, X.; Song, S.Q.; Zhu, M.S.; Lou, Z.Z.; Li, Z.; Liu, F.; Liu, Y.H.; et al. Photocatalytic reduction elimination of UO22+ pollutant under visible light with metal-free sulfur doped g-C3N4 photocatalyst. Appl. Catal. B Environ. 2017, 200, 378–385. [Google Scholar] [CrossRef]
  98. Wang, K.; Li, Q.; Liu, B.S.; Cheng, B.; Ho, W.K.; Yu, J.G. Sulfur-doped g-C3N4 with enhanced photocatalytic CO2-reduction performance. Appl. Catal. B Environ. 2015, 176, 44–52. [Google Scholar] [CrossRef]
  99. Praus, P.; Smykalova, A.; Foniok, K.; Velisek, P.; Cvejn, D.; Zadny, J.; Storch, J. Post-Synthetic Derivatization of Graphitic Carbon Nitride with Methanesulfonyl Chloride: Synthesis, Characterization and Photocatalysis. Nanomaterials 2020, 10, 193. [Google Scholar] [CrossRef] [Green Version]
  100. Qin, H.F.; Lv, W.H.; Bai, J.R.; Zhou, Y.; Wen, Y.P.; He, Q.T.; Tang, J.H.; Wang, L.B.; Zhou, Q.F. Sulfur-doped porous graphitic carbon nitride heterojunction hybrids for enhanced photocatalytic H2 evolution. J. Mater. Sci. 2019, 54, 4811–4820. [Google Scholar] [CrossRef]
  101. Ke, L.; Li, P.F.; Wu, X.; Jiang, S.J.; Luo, M.B.; Liu, Y.H.; Le, Z.G.; Sun, C.Z.; Song, S.Q. Graphene-like sulfur-doped g-C3N4 for photocatalytic reduction elimination of UO22+ under visible Light. Appl. Catal. B Environ. 2017, 205, 319–326. [Google Scholar] [CrossRef]
  102. Raziq, F.; Humayun, M.; Ali, A.; Wang, T.T.; Khan, A.; Fu, Q.Y.; Luo, W.; Zeng, H.P.; Zheng, Z.P.; Khan, B.; et al. Synthesis of S-Doped porous g-C3N4 by using ionic liquids and subsequently coupled with Au-TiO2 for exceptional cocatalyst-free visible-light catalytic activities. Appl. Catal. B Environ. 2018, 237, 1082–1090. [Google Scholar] [CrossRef]
  103. Li, Y.F.; Wang, S.; Chang, W.; Zhang, L.H.; Wu, Z.S.; Song, S.Y.; Xing, Y. Preparation and enhanced photocatalytic performance of sulfur doped terminal-methylated g-C3N4 nanosheets with extended visible-light response. J. Mater. Chem. A 2019, 7, 20640–20648. [Google Scholar] [CrossRef]
  104. Cui, Y.J.; Li, M.; Wang, H.; Yang, C.F.; Meng, S.G.; Chen, F.Y. In-situ synthesis of sulfur doped carbon nitride microsphere for outstanding visible light photocatalytic Cr (VI) reduction. Sep. Purif. Technol. 2018, 199, 251–259. [Google Scholar] [CrossRef]
  105. Mohammad, A.; Khan, M.E.; Cho, M.H. Sulfur-doped-graphitic-carbon nitride (S-g-C3N4) for low cost electrochemical sensing of hydrazine. J. Alloy. Compd. 2020, 816, 10. [Google Scholar] [CrossRef]
  106. Guo, H.; Shu, Z.; Chen, D.H.; Tan, Y.G.; Zhou, J.; Meng, F.Y.; Li, T.T. One-step synthesis of S-doped g-C3N4 nanosheets for improved visible-light photocatalytic hydrogen evolution. Chem. Phys. 2020, 533, 7. [Google Scholar] [CrossRef]
  107. Xie, L.L.; Dai, Y.R.; Zhou, Y.J.; Chang, X.; Yin, L.F. Sulfur (VI) modified graphite carbon nitride nanosheets with chrysanthemum-like structure and enhanced photocatalytic activity. Chem. Phys. Lett. 2018, 693, 1–7. [Google Scholar] [CrossRef]
  108. Cao, J.J.; Wang, H.; Zhao, Y.J.; Liu, Y.; Wu, Q.Y.; Huang, H.; Shao, M.W.; Kang, Z.H. Phosphorus-doped porous carbon nitride for efficient sole production of hydrogen peroxide via photocatalytic water splitting with a two-channel pathway. J. Mater. Chem. A 2020, 8, 3701–3707. [Google Scholar] [CrossRef]
  109. Zhu, Y.K.; Li, J.Z.; Cao, J.M.; Lv, C.X.; Huang, G.Q.; Zhang, G.L.; Xu, Y.; Zhang, S.C.; Meng, P.P.; Zhan, T.R.; et al. Phosphorus-doped polymeric carbon nitride nanosheets for enhanced photocatalytic hydrogen production. APL Mater. 2020, 8, 7. [Google Scholar] [CrossRef] [Green Version]
  110. Chai, B.; Yan, J.T.; Wang, C.L.; Ren, Z.D.; Zhu, Y.C. Enhanced visible light photocatalytic degradation of Rhodamine B over phosphorus doped graphitic carbon nitride. Appl. Surf. Sci. 2017, 391, 376–383. [Google Scholar] [CrossRef]
  111. Chegeni, M.; Dehghan, N. Preparation of Phosphorus Doped Graphitic Carbon Nitride Using a Simple Method and Its Application for Removing Methylene Blue. Phys. Chem. Res. 2020, 8, 31–44. [Google Scholar] [CrossRef]
  112. Huang, J.X.; Li, D.G.; Li, R.B.; Zhang, Q.X.; Chen, T.S.; Liu, H.J.; Liu, Y.; Lv, W.Y.; Liu, G.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]
  113. Huang, J.X.; Li, D.G.; Liu, Y.; Li, R.B.; Chen, P.; Liu, H.J.; Lv, W.Y.; Liu, G.G.; Feng, Y.P. Ultrathin Ag2WO4-coated P-doped g-C3N4 nanosheets with remarkable photocatalytic performance for indomethacin degradation. J. Hazard. Mater. 2020, 392, 13. [Google Scholar] [CrossRef] [PubMed]
  114. Kesavan, T.; Partheeban, T.; Vivekanantha, M.; Prabu, N.; Kundu, M.; Premkumar, S.; Umapathy, S.; Vinu, A.; Sasidharan, M. Design of P-Doped Mesoporous Carbon Nitrides as High-Performance Anode Materials for Li-Ion Battery. ACS Appl. Mater. Interfaces 2020, 12, 24007–24018. [Google Scholar] [CrossRef]
  115. Li, J.J.; Tian, C.; Zhao, H.; Mei, J.; Zhang, J.; Yang, S.J. Controllable fabrication of a red phosphorus modified g-C3N4 photocatalyst with strong interfacial binding for the efficient removal of organic pollutants. J. Alloy. Compd. 2019, 810. [Google Scholar] [CrossRef]
  116. Jolliffe, I.T. Principal Component Analysis. In Principal Component Analysis; Springer: New York, NY, USA, 1986; pp. 115–128. [Google Scholar] [CrossRef]
  117. Ran, J.R.; Guo, W.W.; Wang, H.L.; Zhu, B.C.; Yu, J.G.; Qiao, S.Z. Metal-Free 2D/2D Phosphorene/g-C3N4 Van der Waals Heterojunction for Highly Enhanced Visible-Light Photocatalytic H2 Production. Adv. Mater. 2018, 30, 6. [Google Scholar] [CrossRef]
  118. Su, C.Y.; Zhou, Y.Z.; Zhang, L.L.; Yu, X.H.; Gao, S.; Sun, X.J.; Cheng, C.; Liu, Q.Q.; Yang, J. Enhanced n →pi* electron transition of porous P-doped g-C3N4 nanosheets for improved photocatalytic H2 evolution performance. Ceram. Int. 2020, 46, 8444–8451. [Google Scholar] [CrossRef]
  119. Sun, Y.J.; He, J.Y.; Zhang, D.; Wang, X.J.; Zhao, J.; Liu, R.H.; Li, F.T. Simultaneous construction of dual-site phosphorus modified g-C3N4 and its synergistic mechanism for enhanced visible-light photocatalytic hydrogen evolution. Appl. Surf. Sci. 2020, 517, 8. [Google Scholar] [CrossRef]
  120. Wang, P.Y.; Guo, C.S.; Hou, S.; Zhao, X.; Wu, L.L.; Pei, Y.Y.; Zhang, Y.; Gao, J.F.; Xu, J. Template-free synthesis of bubble-like phosphorus-doped carbon nitride with enhanced visible-light photocatalytic activity. J. Alloy. Compd. 2018, 769, 503–511. [Google Scholar] [CrossRef]
  121. Zhao, Z.L.; Xie, C.; Cui, H.D.; Wang, Q.Y.; Shu, Z.; Zhou, J.; Li, T.T. Scalable one-pot synthesis of phosphorus-doped g-C3N4 nanosheets for enhanced visible-light photocatalytic hydrogen evolution. Diam. Relat. Mat. 2020, 104, 8. [Google Scholar] [CrossRef]
  122. Deng, Y.C.; Tang, L.; Zeng, G.M.; Zhu, Z.J.; Yan, M.; Zhou, Y.Y.; Wang, J.J.; Liu, Y.N. Insight into highly efficient simultaneous photocatalytic removal of Cr (VI) and 2,4-diclorophenol under visible light irradiation by phosphorus doped porous ultrathin g-C3N4 nanosheets from aqueous media: Performance and reaction mechanism. Appl. Catal. B Environ. 2017, 203, 343–354. [Google Scholar] [CrossRef]
  123. Zhu, Y.P.; Ren, T.Z.; Yuan, Z.Y. Mesoporous Phosphorus-Doped g-C3N4 Nanostructured Flowers with Superior Photocatalytic Hydrogen Evolution Performance. ACS Appl. Mater. Interfaces 2015, 7, 16850–16856. [Google Scholar] [CrossRef] [PubMed]
  124. He, F.; Wang, Z.X.; Li, Y.X.; Peng, S.Q.; Liu, B. The nonmetal modulation of composition and morphology of g-C3N4-based photocatalysts. Appl. Catal. B Environ. 2020, 269, 21. [Google Scholar] [CrossRef]
  125. Guo, S.E.; Tang, Y.Q.; Xie, Y.; Tian, C.G.; Feng, Q.M.; Zhou, W.; Jiang, B.J. P-doped tubular g-C3N4 with surface carbon defects: Universal synthesis and enhanced visible-light photocatalytic hydrogen production. Appl. Catal. B Environ. 2017, 218, 664–671. [Google Scholar] [CrossRef]
  126. Zhou, Y.J.; Zhang, L.X.; Liu, J.J.; Fan, X.Q.; Wang, B.Z.; Wang, M.; Ren, W.C.; Wang, J.; Li, M.L.; Shi, J.L. Brand new P-doped g-C3N4: Enhanced photocatalytic activity for H2 evolution and Rhodamine B degradation under visible light. J. Mater. Chem. A 2015, 3, 3862–3867. [Google Scholar] [CrossRef]
  127. Li, X.H.; Zhang, J.; Zhou, F.; Zhang, H.L.; Bai, J.; Wang, Y.J.; Wang, H.Y. Preparation of N-vacancy-doped g-C3N4 with outstanding photocatalytic H2O2 production ability by dielectric barrier discharge plasma treatment. Chin. J. Catal. 2018, 39, 1090–1098. [Google Scholar] [CrossRef]
  128. Li, H.L.; Jin, C.; Wang, Z.Y.; Liu, Y.Y.; Wang, P.; Zheng, Z.K.; Whangbo, M.H.; Kou, L.Z.; Li, Y.J.; Dai, Y.; et al. Effect of the intra- and inter-triazine N-vacancies on the photocatalytic hydrogen evolution of graphitic carbon nitride. Chem. Eng. J. 2019, 369, 263–271. [Google Scholar] [CrossRef]
  129. Peng, G.M.; Wu, J.W.; Wang, M.Z.; Niklas, J.; Zhou, H.; Liu, C. Nitrogen-Defective Polymeric Carbon Nitride Nanolayer Enabled Efficient Electrocatalytic Nitrogen Reduction with High Faradaic Efficiency. Nano Lett. 2020, 20, 2879–2885. [Google Scholar] [CrossRef]
  130. Li, X.X.; Tan, T.N.; Zhang, J.; Zhang, N.; He, J.W.; Han, W.M.; Wang, Y.D.; Pan, M. Nitrogen Deficient Graphitic Carbon Nitride as Anodes for Lithium-ion Batteries. J. Wuhan Univ. Technol. Mat. Sci. Edit. 2020, 35, 263–271. [Google Scholar] [CrossRef]
  131. Xie, Y.; Li, Y.X.; Huang, Z.H.; Zhang, J.Y.; Jia, X.F.; Wang, X.S.; Ye, J.H. Two types of cooperative nitrogen vacancies in polymeric carbon nitride for efficient solar-driven H2O2 evolution. Appl. Catal. B Environ. 2020, 265, 7. [Google Scholar] [CrossRef]
  132. Wang, J.; Gao, B.; Dou, M.M.; Huang, X.; Ma, Z.K. A porous g-C3N4 nanosheets containing nitrogen defects for enhanced photocatalytic removal meropenem: Mechanism, degradation pathway and DFT calculation. Environ. Res. 2020, 184, 10. [Google Scholar] [CrossRef]
  133. Wang, Y.X.; Rao, L.; Wang, P.F.; Guo, Y.; Shi, Z.Y.; Guo, X.; Zhang, L.X. Synthesis of nitrogen vacancies g-C3N4 with increased crystallinity under the controlling of oxalyl dihydrazide: Visible-light-driven photocatalytic activity. Appl. Surf. Sci. 2020, 505, 10. [Google Scholar] [CrossRef]
  134. Yu, W.W.; Shan, X.; Zhao, Z.K. Unique nitrogen-deficient carbon nitride homojunction prepared by a facile inserting-removing strategy as an efficient photocatalyst for visible light-driven hydrogen evolution. Appl. Catal. B Environ. 2020, 269, 9. [Google Scholar] [CrossRef]
  135. Li, X.X.; Zhang, K.L.; Zhou, M.; Yang, K.; Yang, S.; Ma, X.H.; Yu, C.L.; Xie, Y.; Huang, W.Y.; Fan, Q.Z. A Novel Approach to Synthesize Nitrogen-Deficient g-C3N4 for the Enhanced Photocatalytic Contaminant Degradation and Electrocatalytic Hydrogen Evolution. Nano 2020, 15, 13. [Google Scholar] [CrossRef] [Green Version]
  136. Yang, Z.X.; Chu, D.L.; Jia, G.R.; Yao, M.G.; Liu, B.B. Significantly narrowed bandgap and enhanced charge separation in porous, nitrogen-vacancy red g-C3N4 for visible light photocatalytic H2 production. Appl. Surf. Sci. 2020, 504, 8. [Google Scholar] [CrossRef]
  137. Huang, J.J.; Du, J.M.; Du, H.W.; Xu, G.S.; Yuan, Y.P. Control of Nitrogen Vacancy in g-C3N4 by Heat Treatment in an Ammonia Atmosphere for Enhanced Photocatalytic Hydrogen Generation. Acta Phys. Chim. Sin. 2020, 36, 8. [Google Scholar] [CrossRef]
  138. Li, W.S.; Guo, Z.; Jiang, L.T.; Zhong, L.; Li, G.N.; Zhang, J.J.; Fan, K.; Gonzalez-Cortes, S.; Jin, K.J.; Xu, C.J.; et al. Facile in situ reductive synthesis of both nitrogen deficient and protonated g-C3N4 nanosheets for the synergistic enhancement of visible-light H2 evolution. Chem. Sci. 2020, 11, 2716–2728. [Google Scholar] [CrossRef] [Green Version]
  139. Zhang, Y.; Di, J.; Ding, P.H.; Zhao, J.Z.; Gu, K.Z.; Chen, X.L.; Yan, C.; Yin, S.; Xia, J.X.; Li, H.M. Ultrathin g-C3N4 with enriched surface carbon vacancies enables highly efficient photocatalytic nitrogen fixation. J. Colloid Interface Sci. 2019, 553, 530–539. [Google Scholar] [CrossRef]
  140. Li, Y.H.; Gu, M.L.; Shi, T.; Cui, W.; Zhang, X.M.; Dong, F.; Cheng, J.S.; Fan, J.J.; Lv, K.L. Carbon vacancy in C3N4 nanotube: Electronic structure, photocatalysis mechanism and highly enhanced activity. Appl. Catal. B Environ. 2020, 262, 11. [Google Scholar] [CrossRef]
  141. Li, S.N.; Dong, G.H.; Hailili, R.; Yang, L.P.; Li, Y.X.; Wang, F.; Zeng, Y.B.; Wang, C.Y. Effective photocatalytic H2O2 production under visible light irradiation at g-C3N4 modulated by carbon vacancies. Appl. Catal. B Environ. 2016, 190, 26–35. [Google Scholar] [CrossRef] [Green Version]
  142. Barrio, J.; Volokh, M.; Shalom, M. Polymeric carbon nitrides and related metal-free materials for energy and environmental applications. J. Mater. Chem. A 2020, 8, 11075–11116. [Google Scholar] [CrossRef]
  143. Liu, X.L.; Ma, R.; Zhuang, L.; Hu, B.W.; Chen, J.R.; Liu, X.Y.; Wang, X.K. Recent developments of doped g-C3N4 photocatalysts for the degradation of organic pollutants. Crit. Rev. Environ. Sci. Technol. 2020, 40. [Google Scholar] [CrossRef]
  144. Mun, S.J.; Park, S.J. Graphitic Carbon Nitride Materials for Photocatalytic Hydrogen Production via Water Splitting: A Short Review. Catalysts 2019, 9, 805. [Google Scholar] [CrossRef] [Green Version]
  145. Talapaneni, S.N.; Singh, G.; Kim, I.Y.; AlBahily, K.; Al-Muhtaseb, A.H.; Karakoti, A.S.; Tavakkoli, E.; Vinu, A. Nanostructured Carbon Nitrides for CO2 Capture and Conversion. Adv. Mater. 2020, 32, 21. [Google Scholar] [CrossRef]
  146. Wang, Y.Q.; Shen, S.H. Progress and Prospects of Non-Metal Doped Graphitic Carbon Nitride for Improved Photocatalytic Performances. Acta Phys. Chim. Sin. 2020, 36, 14. [Google Scholar] [CrossRef]
  147. Zhu, A.H.; Qiu, B.C.; Du, M.M.; Ji, J.H.; Nasir, M.; Xing, M.Y.; Zhang, J.L. Dopant-Induced Edge and Basal Plane Catalytic Sites on Ultrathin C3N4 Nanosheets for Photocatalytic Water Reduction. Acs Sustain. Chem. Eng. 2020, 8, 7497–7502. [Google Scholar] [CrossRef]
  148. Wanninayake, N.; Ai, Q.; Zhou, R.; Hoque, M.A.; Herrell, S.; Guzman, M.I.; Risko, C.; Kim, D.Y. Understanding the effect of host structure of nitrogen doped ultrananocrystalline diamond electrode on electrochemical carbon dioxide reduction. Carbon 2020, 157, 408–419. [Google Scholar] [CrossRef]
  149. Zhao, H.L.; Liu, L.J.; Andino, J.M.; Li, Y. Bicrystalline TiO2 with controllable anatase-brookite phase content for enhanced CO2 photoreduction to fuels. J. Mater. Chem. A 2013, 1, 8209–8216. [Google Scholar] [CrossRef] [Green Version]
  150. Hoque, M.A.; Guzman, M.I. Photocatalytic Activity: Experimental Features to Report in Heterogeneous Photocatalysis. Materials 2018, 11, 1990. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Mott–Schottky plots for as-synthesized g-C3N4 and O-g-C3N4 samples (a) and postulated band structure of pristine g-C3N4 and increasing oxidation of g-C3N4 (b) Reprinted from [24], Copyright (2020), with permission from Elsevier.
Figure 1. Mott–Schottky plots for as-synthesized g-C3N4 and O-g-C3N4 samples (a) and postulated band structure of pristine g-C3N4 and increasing oxidation of g-C3N4 (b) Reprinted from [24], Copyright (2020), with permission from Elsevier.
Catalysts 10 01119 g001
Figure 2. Band structure transformation with increasing oxygen doping level in g-C3N4. Reprinted from [24], Copyright (2020), with permission from Elsevier.
Figure 2. Band structure transformation with increasing oxygen doping level in g-C3N4. Reprinted from [24], Copyright (2020), with permission from Elsevier.
Catalysts 10 01119 g002
Figure 3. Formation and characterization of g-C3N4/NH3. Schematic illustration of the synthetic process (a), SEM images (b,c), TEM images (d,e), scanning tunneling electron microscopy (STEM) image (f), and energy dispersive X-ray (EDX) maps of the g-C3N4/NH3 nanotubes (TCN (NH3)) (g,h). Reprinted from [62], Copyright (2019), with permission from Elsevier.
Figure 3. Formation and characterization of g-C3N4/NH3. Schematic illustration of the synthetic process (a), SEM images (b,c), TEM images (d,e), scanning tunneling electron microscopy (STEM) image (f), and energy dispersive X-ray (EDX) maps of the g-C3N4/NH3 nanotubes (TCN (NH3)) (g,h). Reprinted from [62], Copyright (2019), with permission from Elsevier.
Catalysts 10 01119 g003
Figure 4. Scheme of the formation process from the bulk-C3N4 to the g-C3N4 nanosheets (NCNS) sample. Reprinted from [57], Copyright (2018), with permission from Elsevier.
Figure 4. Scheme of the formation process from the bulk-C3N4 to the g-C3N4 nanosheets (NCNS) sample. Reprinted from [57], Copyright (2018), with permission from Elsevier.
Catalysts 10 01119 g004
Figure 5. Ultra-violet diffuse-reflectance spectra of bulk g-C3N4, C/CNNS-x (x = 0.1, 0.3, 0.4, 0.5, 1.0, 2.0) (a), The band gap plots (b), valence band (VB) XPS spectra (c), and the band structure diagrams of bulk g-C3N4 and C/CNNS-0.5 photocatalysts (d). Reprinted from [79], Copyright (2019), with permission from Elsevier.
Figure 5. Ultra-violet diffuse-reflectance spectra of bulk g-C3N4, C/CNNS-x (x = 0.1, 0.3, 0.4, 0.5, 1.0, 2.0) (a), The band gap plots (b), valence band (VB) XPS spectra (c), and the band structure diagrams of bulk g-C3N4 and C/CNNS-0.5 photocatalysts (d). Reprinted from [79], Copyright (2019), with permission from Elsevier.
Catalysts 10 01119 g005
Figure 6. Optical absorption spectra of g-C3N4 and S-g-C3N4 (SC3N4) samples (a); plots of (αhυ)1/2 vs. photon energy (hυ) (b) and electrochemical Mott–Schottky plots of g-C3N4 and S-g-C3N4 samples (c). Reprinted from [101] Copyright (2017), with permission from Elsevier.
Figure 6. Optical absorption spectra of g-C3N4 and S-g-C3N4 (SC3N4) samples (a); plots of (αhυ)1/2 vs. photon energy (hυ) (b) and electrochemical Mott–Schottky plots of g-C3N4 and S-g-C3N4 samples (c). Reprinted from [101] Copyright (2017), with permission from Elsevier.
Catalysts 10 01119 g006
Figure 7. Scheme of the formation of melamine-trithiocyanuric acid (MT) complex and S-g-C3N4. Reprinted from [95] Copyright (2017), with permission from Elsevier.
Figure 7. Scheme of the formation of melamine-trithiocyanuric acid (MT) complex and S-g-C3N4. Reprinted from [95] Copyright (2017), with permission from Elsevier.
Catalysts 10 01119 g007
Figure 8. Scheme of a possible existing form of P atoms in the framework of of g-C3N4 and its effect on the electronic and chemical properties of g-C3N4. The dark blue P atom refers to the corner site and the red one is the bay site. Reproduced from [126] Copyright (2015), with permission from The Royal Society of Chemistry.
Figure 8. Scheme of a possible existing form of P atoms in the framework of of g-C3N4 and its effect on the electronic and chemical properties of g-C3N4. The dark blue P atom refers to the corner site and the red one is the bay site. Reproduced from [126] Copyright (2015), with permission from The Royal Society of Chemistry.
Catalysts 10 01119 g008
Figure 9. 13C solid-state magic-angle spinning NMR spectra (a), electron paramagnetic resonance (EPR) spectra of g-C3N4 (CN) and oxalyl dihydrazide (ODH)-CN2 (b), structure model of ODH-CN with N vacancies (c). Reprinted from [133] Copyright (2020), with permission from Elsevier.
Figure 9. 13C solid-state magic-angle spinning NMR spectra (a), electron paramagnetic resonance (EPR) spectra of g-C3N4 (CN) and oxalyl dihydrazide (ODH)-CN2 (b), structure model of ODH-CN with N vacancies (c). Reprinted from [133] Copyright (2020), with permission from Elsevier.
Catalysts 10 01119 g009
Figure 10. Schematic illustration of CO2 capture and conversion to organic fuels using carbon nitride nanostructures (a), molecular structure (b) and band structure (c) of g-C3N4. Reproduced with permission [145], Copyright 2019, with permission from John Wiley and Sons.
Figure 10. Schematic illustration of CO2 capture and conversion to organic fuels using carbon nitride nanostructures (a), molecular structure (b) and band structure (c) of g-C3N4. Reproduced with permission [145], Copyright 2019, with permission from John Wiley and Sons.
Catalysts 10 01119 g010
Figure 11. Proposed mechanism for Pt-deposited g-C3N4 (CN) and C-g-C3N4 (CD-CN) of visible-light photocatalytic performance; TEOA is triethanolamine and D is oxidation products. Reprinted from [71] Copyright (2020), with permission from Elsevier.
Figure 11. Proposed mechanism for Pt-deposited g-C3N4 (CN) and C-g-C3N4 (CD-CN) of visible-light photocatalytic performance; TEOA is triethanolamine and D is oxidation products. Reprinted from [71] Copyright (2020), with permission from Elsevier.
Catalysts 10 01119 g011
Figure 12. The scheme of the RhB photocatalytic degradation in the presence of O-g-C3N4. Reprinted from [17].
Figure 12. The scheme of the RhB photocatalytic degradation in the presence of O-g-C3N4. Reprinted from [17].
Catalysts 10 01119 g012
Table 1. Synthetic methods, applications, and photocatalytic efficiency of O-doped g-C3N4.
Table 1. Synthetic methods, applications, and photocatalytic efficiency of O-doped g-C3N4.
PrecursorSynthetic MethodC/O Ratio,
Doped (Pristine)
Photocatalytic ProcessConditions of the ProcessEfficiency
Doped/Pristine
References
UreaH2O2 hydrothermal treatment, 120 °CSurface O at.% in O-doped g-C3N4
3.23–6.59 by XPS
H2 evolution500 W Xe lamp, simulate solar irradiation; 30 vol% triethanolamine (TEOA)408.4 μmol·g−1/
317.9 μmol·g−1
[24]
MelaminePMS hydrothermal treatment, 60 °C3.16%/2.73%
by Energy-Dispersive X-Ray Spectroscopy (EDS)
1.8%/2.8% by XPS
Rhodamine B (RhB) destruction500 W halogen lamp,
filter λ > 420 nm
0.079 min−1/0.0032 min−1[17]
UreaH2O2 hydrothermal treatment, 120 °CRhB, Methyl orange (MO) destruction500 W Xe lamp, simulated solar irradiation RhB:0.1074 h−1/0.0170 h−1
MO: 0.2287 h−1/0.0095 h−1
[41]
Urea, ammonium acetatetwo-step thermal treatment4.3/8.8 by elemental analysis (EA)bisphenol A (BPA), phenol (Ph), 2-chlorphenol (2-Ph), diphenhydramine destruction (DP)Light-emitting diode (LED) lamp,
λ = 420–780 nm
Total organic carbon (TOC) removal rate:
BPA 72.79%/9.27%;
Ph 67.3%, 2-Ph 61.5%, DP 55.0%
[19]
Melamine, ethanolThermal
polymerization
Atom.% (O)
3.25/- by XPS
H2 evolution350 W Xe lamp, filter λ > 420 nm,
10% vol% TEOA, Pt co-catalyst (1 wt.%)
64.30 μmol·h−1/
3.6 μmol·h−1
[42]
1,3,5-Trichloro-triazine, dicyandiami-deSolvothermal
method, 200 °C
C:N:O
1.08:1:0.23/
0.69:1:0.03
by XPS
H2 evolution;
RhB destruction
Visible light, λ > 420, Pt co-catalyst (0.3 wt.%)H2 evolution:
3174 μmol·h−1 g −1/846 μmol·h−1·g −1
RhB degradation:
0.249 min−1/0.007min−1
[22]
MelamineThermal
polycondensation
12.5/trace by XPSCO2 reduction350 W Xe lamp, filter λ > 420 nmCH3OH production
0.88 µmol·g−1·h−1/
0.17 µmol·g−1·h−1
[43]
Dicyandia-midineHydrothermal followed by calcinationO content (wt.%)
5.23/0.17 by EA
N2 fixation500 W Xenon lamp filter λ > 420 nm,
10 vol.% methanol as sacrificial agents
118.8 mg·l−1·h−1·gcat−1/
5.86 mg·l−1·h−1·gcat−1
[18]
Table 3. Synthetic methods, applications, and photocatalytic efficiency of C-doped g-C3N4.
Table 3. Synthetic methods, applications, and photocatalytic efficiency of C-doped g-C3N4.
PrecursorSynthetic MethodC/N
Doped (Pristine)
Photocatalytic ProcessConditions of the ProcessEfficiency Doped/PristineReferences
Dicyanamide, dimethylformamideThermal copolymerizationC/N mass ratio
0.61(0.59) by elemental analysis
H2 evolution300W Xe lamp, filter λ > 400 nm), Pt co-catalyst (1 wt.%), triethanolamine (TEOA) as a hole quencher35.5 μmol/
6.78 μmol in 8 h
[71]
Melamine,
cellulose
thermal treatmentC/N mass ratio
33.39 (30.12) by elemental analysis
H2 evolution300W Xe lamp, filter λ > 420 nm), Pt (3%), TEOA (10 vol%)1024 μmol·g−1·h−1/
59.6 μmol·g−1·h−1
[68]
Melamine,
Urea, phenylmalonic acid
Precursor copolymerization on the surface of g-C3N4C/N at. Ratio
68.9 (43.0)
C, %
33.52 (33.45) by elemental analysis
bisphenol A
(BPA) destruction
H2 evolution
300W Xe lamp, filter λ > 420 nm, Pt (3%), TEOA (10 vol%)BPA destruction:
0.0507 min−1/0.0038 min−1;
H2 evolution:
31 μmol·h−1/10 μmol·h−1
[73]
Cyanuric acid, ethylene glycol, melaminemicrowave treatment of supramolecular aggregatesC/N at. Ratio
0.688 (0.669)
C, %:
39.98 (39.51) by elemental analysis
N2 photofixation250W high-pressure Na lamp (400 < l < 800 nm) 5.3 mg·L−1·gcat−1/
0.48 mg·L−1·gcat−1
[88]
Urea, C60 nanorodsliquid-liquid
interfacial precipitation method
-H2 evolution500 W Xe lamp, filter λ > 420 nm, Pt (3%), TEOA (17 vol%)8.7 μmol·h−1/
1.85 μmol·h−1
[76]
Agar-melamine gelone-step thermal condensation methodC/N at. Ratio
0.69 (0.67)
C, %:
34.9 (35.33) by elemental analysis
C/N at. Ratio
0.90 (0.87) by XPS
RhB, Phenol, BPA, Phe destruction300 W Xe lamp, filter λ > 420 nmRhB destruction
0.042 min−1/
0.016 min−1
BPA destruction
0.145 min−1/
0.113min−1
[79]
Urea, sacharoseThermal polymerizationC/N at. Ratio
0.58 (0.57)
C, %:
34.83 (34.81) by XPS analysis
NO removalXe lamp, filter λ > 420 nmNO removal ratio 56.77%/
50.89%
[82]
Melamine, carbon dots (CD)combining g-C3N4 treated with H2O2 and CDC wt.%:
46.77(39.78) by EDX
MB, RhB, fuchsine, Phe destruction; Cr(VI) photoreduction50 W LED lamp, visible light irradiationRhB destruction
0.0675 min−1/
0.0019 min−1
[67]
Table 4. Synthetic methods, applications, and photocatalytic efficiency of S-doped g-C3N4.
Table 4. Synthetic methods, applications, and photocatalytic efficiency of S-doped g-C3N4.
PrecursorSynthetic MethodContent of SulfurPhotocatalytic ProcessConditions of the ProcessEfficiency Doped/PristineReferences
Melamine, sulfurThermal
polycondensation, 520 °C
-O2 evolution;
scheme H2O splitting;
bactericidal activity
400 W halide lamp (λmax = 360 nm), 150 W Xe lamp, filter λ > 400 nm
[Co(bpy)3]SO4 as electron mediator, Ru/SrTiO3:Rh as cocatalyst
O2 evolution:
40.3 μmol·h−1/-
Z-scheme H2O splitting:
29.3 μmol·h−1/-;
70% of bacteria were killed
[96]
Melamine,
(NH4)2SO4
Upgraded gas templating methodH2 evolution350 W Xe lamp, filter λ > 420 nm, 10% vol% TEOA, Pt co-catalyst (1 wt.%)572 μmol·h−1·g−1/
78 μmol·h−1·g−1
[106]
Urea,
thioacetamide
One-pot copolymerization0.1 at.% S
0.2 by XPS
Procion Red MX-5B degradation
H2 evolution
500 W Xe lamp, monochromic light provided by using a 420 ± 15, 450 ± 15, 475 ± 15 and 520 ± 15 nm band pass filter
TEOA, Pt co-catalyst (1 wt.%)
Dye degradation
0.072 min−1/0.024 min−1
H2 evolution
3.17 mmol·g−1·h−1/
0.84 mmol·g−1·h−1
[103]
ThioureaThermal
polycondensation followed by thermal oxidative etching
S content: 0.45 by OEA, 1.58 by XPSPhenol degradation;
H2 evolution
300 W Xe lamp, 5% vol% TEOA, Pt co-catalyst (1 wt.%)75%/100% of Phenol was decomposed;
H2 evolution:
127.4 μmol·h−1/
0.5 μmol·h−1
[90]
Thiourea, mesyl chloride Post-synthetic derivatization of
g-C3N4
-Acid Orange 7 dye degradationUVA tube lamp, λmax = 368 nm0.113 min−1/0.022 min−1[99]
1,3,5-trichlorotriazine,
Melamine,
Trithiocyanuric acid
Solvothermal condensation process, 180 °C-Cr(VI) reductionIrradiation with λ > 420 nm 1.85 min−1/0.03 min−1[104]
Melamine,
Trithiocyanuric acid
Thermal
polycondensation of the supramolecular complex
S (wt%): 0.63RhB degradation500 W Xe lamp, filter λ > 420 nm0.0167 min−1/0.0013 min−1[95]
Urea, benzyl disulfideThermal
polycondensation, 520 °C
-Reduction elimination of UO22+350 W Xe lamp, λ ≥ 420 nm0.16 min−1/0.07 min−1[101]
Urea, thiourea Thermal
polycondensation, 550 °C
-H2 evolution300 W Xe lamp, filter λ > 420 nm, 10% vol% TEOA95.3 μmol·h−1/
36.4 μmol·h−1
[100]
Melamine, sulfuric acidSulfuring and sonicating bulk g-C3N4-4-nitrophenol degradation 500 W Xe lamp, filter λ > 400 nm3.47 × 10−2 min−1/
7.04 × 10−4 min−1
[107]
ThioureaThermal
polycondensation, 520 °C
S atomic% 0.05 by EACO2 reduction300 W Xe lamp, Pt co-catalyst (1 wt.%)CH3OH formation:
1.12 μmol·g−1/0.81 μmol·g−1
[98]
Table 6. Synthetic methods, applications, and photocatalytic efficiency of vacancy-doped g-C3N4.
Table 6. Synthetic methods, applications, and photocatalytic efficiency of vacancy-doped g-C3N4.
PrecursorSynthetic MethodC/N Element Atomic Ratio,
Doped (Pristine)
Photocatalytic ProcessConditions of the ProcessEfficiency Doped/PristineReferences
Urea,
oxalyl dihydrazide (ODH)
Thermal copolymerization0.74 (0.65) by element analysis;
0.67 (0.61) by XPS
1.87 (1.09) by EDX
Tetracycline hydrochloride (TC-HCl) and sulfamethoxazole (SMZ) destruction;
H2 evolution
300 W Xe lamp,
filter λ > 420 nm,
Pt co-catalyst (1wt.%), TEOA
SMZ destruction:
0.0203min−1/
0.0066 min−1;
H2 evolution:
5833.1 μmol·h−1·g−1/
1458.2 μmol·g−1
[133]
Urea, dicyandiamidePost-thermal treatment of
g-C3N4
H2 evolutionVisible light irradiation6.5 μmol·g−1/
2.1 μmol·g−1
[137]
MelaminePolymerization in atmosphere of: CCl4; H2; Ar0.61 (0.65) by elemental analysisH2 evolution300W Xe lamp, filter λ > 420 nm), Pt (3%), TEOA (10 vol%)0.079 min−1/
0.0032 min−1
[128]
UreaKOH-assisted calcination treatment1.45 (1.51) by organic elemental analysis
1.32 (1.64) by XPS
H2O2 productionsimulated sunlight lamp
20 vol% ethanol
152.6 μmol·h−1/
10.2 μmol·h−1
[131]
Dicyandiamide, NH4Cl, 3-amino-1,2,4-triazolThermal polymerization with post treatment in N21.260 (1.489) by element analysisH2 evolution300W Xe lamp, filter λ > 400 nm), Pt (3%), TEOA (10 vol%)3882.5 μmol·h−1·g−1/
85.0 μmol·h−1·g−1
[134]
Urea, Mg powdermagnesium vapor etching0.51 (0.78) by EDX
0.92 (1.14) by XPS
H2 evolution300W Xe lamp, filter λ > 400 nm), Pt (3%), TEOA (10 vol%)450 μmol·h−1·g−1/
225 μmol·h−1·g−1
[88]
Dicyandiamidetwo-step calcination0.81 (0.85) by XPSN2 fixation 300W Xe lampNH4+ formation:
54 mmol·L−1/
24 mmol·L−1
[139]
Urea, melamineprecursor preprocessing and thermolysis in N2NO oxidationLED lamp (λ ≥ 448 nm)the NO oxidation in 30 min of irradiation
47.7%/22%
[140]

Share and Cite

MDPI and ACS Style

Starukh, H.; Praus, P. Doping of Graphitic Carbon Nitride with Non-Metal Elements and Its Applications in Photocatalysis. Catalysts 2020, 10, 1119. https://doi.org/10.3390/catal10101119

AMA Style

Starukh H, Praus P. Doping of Graphitic Carbon Nitride with Non-Metal Elements and Its Applications in Photocatalysis. Catalysts. 2020; 10(10):1119. https://doi.org/10.3390/catal10101119

Chicago/Turabian Style

Starukh, Halyna, and Petr Praus. 2020. "Doping of Graphitic Carbon Nitride with Non-Metal Elements and Its Applications in Photocatalysis" Catalysts 10, no. 10: 1119. https://doi.org/10.3390/catal10101119

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