Next Article in Journal
Ammonium-Salt Formation and Catalyst Deactivation in the SCR System for a Marine Diesel Engine
Next Article in Special Issue
CdS Photocatalysts Modified with Ag: Effect of the Solvothermal Temperature on the Structure and Photoactivity for Hydrogen Production
Previous Article in Journal
Destruction of Toluene, Naphthalene and Phenanthrene as Model Tar Compounds in a Modified Rotating Gliding Arc Discharge Reactor
Previous Article in Special Issue
Synthesis, Characterization of g-C3N4/SrTiO3 Heterojunctions and Photocatalytic Activity for Organic Pollutants Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Development of Photocatalysts Containing Carbon Species: A Review

1
Key Laboratory of Forest Resources Conservation and Utilization in the Southwest Mountains of China, Southwest Forestry University, Ministry of Education, Kunming 650224, China
2
College of Chemical Engineering, Southwest Forestry University, Kunming 650224, China
3
Engineering Research Center of Forestry Biomass Materials and Bioenergy, Beijing Key Laboratory of Lignocellulosic Chemistry, College of Materials Science and Technology, Beijing Forestry University, Beijing 100083, China
4
Department of Chemistry, the Ångström Laboratory, Uppsala University, 75121 Uppsala, Sweden
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(1), 20; https://doi.org/10.3390/catal9010020
Submission received: 3 December 2018 / Revised: 21 December 2018 / Accepted: 24 December 2018 / Published: 28 December 2018
(This article belongs to the Special Issue Photocatalysis Science and Engineering in Europe)

Abstract

:
Undoubtedly, carbon-based (nano)composites can be promising photocatalysts with improved photocatalytic activity due to the coupling effect from the incorporation of carbon species. In this mini-review, we focus on the recent development of photocatalysts based on carbon-based (nano)composites. TiO2 is well-known as a typical photocatalyst. Special attention is paid to the various types of carbon–TiO2 composites such as C-doped TiO2, N–C-doped TiO2, metal–C-doped TiO2, and other co-doped C/TiO2 composites. Various synthetic strategies including the solvothermal/hydrothermal method, sol–gel method, and template-directed method are reviewed for the preparation of carbon-based TiO2 composites. C/graphitic carbon nitride (g-C3N4) composites and ternary C-doped composites are also summarized and ascribed to the unique electronic structure of g-C3N4 and the synergistic effect of the ternary interfaces, respectively. In the end, we put forward the future perspective of the photocatalysts containing carbon species based on our knowledge.

Graphical Abstract

1. Introduction

Photocatalysis has been widely used in many applications such as water splitting [1], energy conversion [2,3,4], and environmental remediation [5]. In recent years, environmental problems have attracted more and more attention. Obviously, photocatalysis is one of potential strategies for the degradation of pollutants, water treatment, and energy utilization [6,7,8]. Various materials including metal oxides [9,10,11], metal sulfides [12,13,14], graphitic carbon nitride (g-C3N4) [15,16,17], and composites [18,19,20] have been reported as photocatalysts. Among these materials, composite photocatalysts have received special attention due to the synergistic effect from different components [21,22,23]. For example, the graphene/TiO2 nanocomposite was reported to improve visible light photocatalytic activity, which displayed efficient visible light photocatalytic applications [24,25,26]. The incorporation of carbon species has been proven to be a promising strategy for improving the photocatalytic efficiency of composites, due to their high specific surface area, unique structure, or other properties [27,28,29]. Therefore, photocatalysts containing carbon species have been hot research materials for potential applications in environmental and energy fields.
This current mini-review gives an overview of the synthesis and photocatalytic properties of photocatalysts containing carbon species, especially C-doped TiO2. In Section 2, various methods including the solvothermal/hydrothermal method, sol–gel method, and template-directed method for the synthesis of C-doped TiO2 are summarized. In addition, C-doped TiO2 with pores such as mesoporous and hollow structures is introduced via typical examples. Finally, the existing problems, future development trends, and potential applications of these promising photocatalysts are put forward. We expect that this review would catch a glimpse of the current developments of these photocatalysts containing carbon species.

2. Photocatalysts of C-doped TiO2

2.1. Photocatalysts of TiO2/C Composites

Titanium dioxide (TiO2) is well-known as a typical photocatalytic metal oxide. However, it displays low photocatalytic efficiency under visible-light irradiation due to its intrinsic large band gap. The incorporation of carbon species into TiO2 for the fabrication of C-doped TiO2 is a promising strategy for improving the photocatalytic efficiency, especially under visible light irradiation. In early 2006, Shanmugam et al. synthesized TiO2@C core–shell composite nanoparticles in the range of 15–35 nm in a specially-made Swagelok cell at different temperatures [30]. They discovered that the TiO2@C composites showed high photocatalytic activity for 4-chlorophenol degradation, compared with Degussa P25. Then, Sullivan et al. also presented visible light active C-doped TiO2 photocatalysts [31], displaying the activities in the oxidation of 4-chlorophenol under visible light irradiation. In their work, through the initial He heat treatments, one can tune the extent of graphitic carbon laydown. After that, C and N single and co-doping TiO2 films of single-layer electrodes were prepared as dye-sensitized solar cell devices [32], enhancing their short circuit current and efficiency, and improving the electron lifetime, charge transport, and recombination resistance. They achieved the highest cell efficiency of 10.2%, as a result of a balance between the band gap, surface area, and diffused reflection of the photoanode. More recently, visible light active carbon-doped TiO2 with anatase/rutile multi-phases coated on granular activated carbon was applied to the removal of nonylphenol in solution [33]. They discovered the decreased band gap from 3.17 eV to 2.72 eV after C doping.
The solvothermal method is one traditional route for the synthesis of metal oxides. In the literature, the calcination-assisted solvothermal method was developed to fabricate the C–TiO2 composites coupled with three colors (blue, green, and red) of emitting up-conversion phosphors [34], displaying the unique optical property via absorbing UV, visible, and near-infrared (NIR) lights. These composites exhibited photocatalytic activities for the degradation of rhodamine B (RhB) solution under NIR light irradiation and the destruction of continuous NO gas flow under the irradiation of UV, visible light-emitting diodes (LEDs), and infrared laser. It was found that the calculated photon efficiency under infrared light irradiation was much higher than those under UV and visible light irradiations, and the green light emitting up-conversion phosphor coupled C–TiO2 composite presented the superior photocatalytic performance over the blue and red ones [34].
The solvothermal method was also applied to prepare TiO2/graphene composites with a chemical bonding interface and enhanced photocatalytic activity in the photodegradation of formaldehyde in air using tetrabutyl orthotitanate as the Ti source [35]. As for the TiO2/graphene nanocomposites with 2.5 wt% graphene, numerous TiO2 nanocrystals that were densely deposited onto the graphene sheets were obtained due to interfacial interactions and preferential heterogeneous nucleation (Figure 1a,b). One can observe the fringe spacing of ca. 3.51 Å matching the (101) crystallographic plane of anatase TiO2 (Figure 1d), exhibiting a high surface energy. Interestingly, in their work, they found the graphene content to exhibit an obvious influence on the photocatalytic activity and synergetic effect between graphene and TiO2 nanoparticles. The authors indicated that the TiO2/graphene composite with chemical bonding caused an intimate interaction between TiO2 nanoparticles and graphene nanosheets (Figure 2a). The formation of a C–Ti bond significantly contributed to the enhanced photocatalytic activity of the TiO2/graphene composite. It is energetically feasible to transfer the photogenerated electrons on the conduction band (CB) of TiO2 to graphene (Figure 2b). It was reported that the energy level of the valence band (VB) is 2.81 V, the energy level of the conduction band (CB) is −0.39 V, and the calculated Fermi energy level of graphene is −0.08 V versus a normal hydrogen electrode, respectively. The authors suggested that the design of new heterogeneous photocatalysts can be applied in environmental protection, water splitting, and photoelectrochemical conversion.
There have been a few reports on the synthesis of C-doped TiO2 composites via the sol-gel method. For example, the sol–gel method was applied to prepare the core–shell structure of TiO2–carbon nanoparticles by utilizing the organic group in the Ti precursor [36]. One can observe the core–shell morphology with a ~1.5-nm thick amorphous carbon layer, acting as a barrier on the grain boundary to obstruct the growth of anatase TiO2, resulting in a decrease in its crystal size of nine nm (Figure 3a–c). The Electron Energy Loss (EELS) spectra showed an average decrease in the Ti valence state in both C/TiO2 and reduced TiO2 (Figure 3d). They reported a large number of oxygen vacancies for these materials, inducing the valence band tails, to enhance visible light absorption. The C/TiO2 showed excellent solar-driven photocatalytic activities on the photodegradation of phenol and methyl blue, and an excellent hydrogen generation rate, which was 10 times higher than that of the reference TiO2 by photosplitting water.
The sol–gel method was also used to synthesize TiO2/C nanosheets with a unique, multi-layer, cross-linked hole structure via the assistance of 1-butyl-3-methylimidazolium tetrafluoroborate ([Bmim][BF4]), which is a polymerization of aniline and carbonization [37]. They discovered that the TiO2/C nanosheets showed an amazing photocatalytic capacity under both UV and sunlight irradiations due to C-doping, as well as a unique morphology, structure, and mixed phases of anatase and rutile, with high crystallization. The sol–gel method was developed to prepare a C-doped TiO2 catalyst for the visible light-induced photocatalytic degradation of ethylene [38]. They evaluated the effects of key parameters such as the visible light intensity, temperature, and feed composition (ethylene, oxygen, and water vapor) on the photocatalytic reaction rate. They indicated that the oxidation rate of ethylene was improved significantly with an increase of the visible light intensity, temperature, and oxygen and ethylene concentrations. Hassan et al. applied a modified sol–gel route to synthesize C–TiO2 thin films based on the self-assembly technique, exploiting Tween80 as a pore directing agent and carbon source [39]. It was reported that the addition of carbon enhanced the photocatalytic activity of TiO2 for methyl orange degradation under visible light irradiation. Recently, an in situ modified sol–gel method was used to synthesize the mesoporous C-doped TiO2 crystals with an average grain size of 22 nm and oxygen vacancy by the limited-oxygen atmosphere originating from the decomposition of urea [40]. These materials displayed 97.3% removal efficiency of RhB in nine minutes under UV light, and they also showed the complete RhB degradation under natural sunlight irradiation.
The porous structure is of great importance for increasing and improving the photoactivity of C-doped TiO2 nanomaterials due to their large surface area and pore volume. In the previous literature, the mesoporous C-doped anatase TiO2 nanomaterials with excellent visible light photocatalytic activity in the degradation of toluene in the gas phase were synthesized by a one-step “green” synthetic approach with low-cost inorganic Ti(SO4)2 and glucose as precursors [41]. They indicated that the new electronic states above the valence band edge were directly responsible for the electronic origin of the band gap narrowing and visible light photoactivity of the C-doped TiO2. After that, mesoporous C-doped anatase TiO2 nanomaterials were also prepared by a one-pot green synthetic approach and post-thermal treatment using sucrose as a carbon-doping source [42]. The enhancement effect of post-thermal treatment between 100–300 °C was proved by the photodegradation of gas-phase toluene. The promoting effect of the post-thermal treatment can be attributed to the changes of the catalysts’ surface and optical properties. The recombination of electron–hole pairs is effectively inhibited after thermal treatment due to the reduction of surface defects.
Zhou et al. synthesized uniformly carbonaceous and nitrogenous species sensitized TiO2 quantum dots employing an oleic acid and oleylamine system and the mesoporous structure of C/N–TiO2 nanospheres with significantly enhanced photocatalytic activity via an emulsion-based bottom–up self-assembly [43]. They obtained the narrow size distribution of monodisperse C/N–TiO2 quantum dots at an average size of 3.8 nm due to the effect of oleylamine and oleic acid, and mesostructured C/N- TiO2 nanospheres ranging from 100 nm to 300 nm in diameter, which were composed of these C/N–TiO2 quantum dots. Moreover, the mesoporous structure of the C/N–TiO2 nanospheres was stable in the degradation reactions irradiated by visible light. By measuring the photodegradation of RhB under visible light irradiation, they reported the water miscible mesostructured C/N–TiO2 nanospheres with improved catalytic area and RhB adsorption, and a red shift of the absorption edge to the visible light region, possessing improved photocatalytic activity.
The template-directed method has been a promising strategy for the synthesis of hollow structures. In Shi’s work, carbon-doped TiO2 hollow spheres with a unique three-dimensional network structure and high visible light photocatalytic activity were synthesized using carbon spheres as the template [44]. By the decoration of methylene blue solution under visible light irradiation, carbon-doped TiO2 hollow spheres were reported to possess higher visible light-induced photocatalytic activity than commercial P25, which could be attributed to their visible light absorption characteristic created by C-doping and the unique hollow three-dimensional network structure. The template-directed self-assembly and calcination method was also explored to create C-doped hollow TiO2 spheres in situ [45]. The hollow and porous structure, as well as a highly ordered and well-defined morphology, were obtained. A number of broken spheres were observed in the sample prepared using one gram of tetrabutyl orthotitanate. The hollow sphere had the uniform shell thickness of ∼20 nm in the sample prepared using two grams of tetrabutyl orthotitanate (Figure 4a,b). Increasing the amount of tetrabutyl orthotitanate to three grams, four grams, and five grams, hollow TiO2 spheres with shell thickness of ∼29 nm (Figure 4c,d), 45 nm (Figure 4e,f), and 59 nm (Figure 4g,h) were obtained, respectively. C-doped hollow TiO2 spheres, as photocatalysts for the degradation of RhB under visible light irradiation, exhibited a superior photocatalytic activity to commercial P25 and pure TiO2-based photocatalysts.
Using a multi-walled carbon nanotube mat as a "rigid" template and carbon-doping source, a C-doped mesoporous anatase TiO2 with a high surface area was obtained [46]. TiO2 nanoparticles have an average diameter of 10 nm, surface area of 112–129 m2 g−1, and total pore volume of 0.74-0.85 m2 g−1. Using banana (Musa acuminata) stem fiber as a sacrificial template, visible light active C-doped mesoporous anatase TiO2 multi-tubes with highly improved visible light photocatalytic activity were synthesized [47]. The as-obtained materials showed a high surface area of 99 m2 g−1, a mesoporous structure made of 15 ± 3 nm nanoparticles, an enhanced light absorption property in the whole visible light region, good thermal stability of the anatase phase up to 750 °C, and an excellent photocatalytic activity for the reduction of Cr(VI) to Cr(III) under the visible light exposure.
Recently, carbon composites derived from different environmentally friendly renewable biomass attracted considerable attention due to their unique properties and wide applications [48,49,50]. Liu et al. fabricated visible light-active C/TiO2 composites by the calcination method using C16H36O4Ti as the titanium source and bamboo powder with different sizes as the carbon source [51]. The as-obtained composites displayed the photocatalytic activities both under visible light irradiation and UV irradiation. After four cycles, bamboo C/TiO2 composites still exhibited high stability. Moreover, they prepared biomass charcoal-doped titanium dioxide (C/TiO2) composites by the microwave-hydrothermal and calcination methods using tetrabutyl titanate as the titanium source and lignin as the carbon source [52]. The composites had a large BET-specific surface area of 203.58 m2 g−1, a total pore volume of 0.16 cm3 g−1, and an average pore diameter of 22.9 nm. Experimental results indicated that both the microwave-hydrothermal temperature and time played an important role in the microstructure and photocatalytic activity. The core–shell structured CdS-decorated TiO2/carbon microspheres were synthesized with controlled hydrolysis and the sonochemical method [53]. The pure commercial TiO2 had only the degradation efficiency of 4.96% for the RhB, and the TiO2/carbon reached 18.92%, while the CdS-decorated TiO2/carbon microspheres achieved 95.24% under visible light irradiation.

2.2. Photocatalysts of N–C-Doped TiO2

In comparison with C-doped TiO2 composites, N–C-doped TiO2 composites have been found to exhibit improved photocatalytic activity. Wang and Lim synthesized a highly visible light photoactive carbon and nitrogen co-doped TiO2 with well-developed mesoporosity and a large BET surface area of 102 m2 g−1 via the solvothermal method [54], and they used these photocatalysts for bisphenol A degradation in a photoreactor irradiated with flexible strips of visible light-emitting diode lights. They reported that under five hours of irradiation with white, blue, green, and yellow light-emitting diode lights, the extents of bisphenol A degradation were >99%, >99%, 84%, and 24%, respectively, while the corresponding percentages of mineralization were 70%, 60%, 45%, and 9%, respectively. The hydrothermal method was developed to synthesize the nitrogen and carbon co-doped titanium dioxide photocatalysts using nitrogen from ammonia and carbon from alcohols with different chain lengths (methanol, ethanol, isopropanol, 1-butanol, 2-butanol, and tertrabutanol) as carbon precursors [55]. The authors observed the increasing activity of phenol decomposition with the chain length of the alcohol precursor. They discovered that the highest intrinsic photoactivity for N and C co-doped TiO2, which was prepared from 2-butanol and 1-butanol as carbon precursors, exhibited a strong resistance to deactivation during multiple catalyst reuse compared with pristine TiO2 and commercial Degussa P25 photocatalysts. In Xu et al.’s work, the hydrothermal treatment was employed to prepare N-doped TiO2/C nanocomposites with high visible light photocatalytic activity in the presence of L-lysine as a ligand to control the nanocrystal growth and a source of nitrogen and carbon [56]. As-prepared nanocomposites had improved absorption in the visible light region and exhibited a higher photocatalytic activity, compared with pure TiO2, commercial P25, and previously reported N-doped TiO2 photocatalysts. The molar ratio of L-lysine/TiCl4 and the pH value were found to be the important factors affecting the photocatalytic activity of the N–TiO2/C composites.
The peptization-reflux method was reported to synthesize visible light-activated C–N–TiO2 with a diameter of ca. 7 nm at 120 °C [57], showing a shift of the absorption edge to a lower energy, a stronger absorption in the visible region, and much higher photocatalytic activity for methyl orange degradation under visible light irradiation, compared to TiO2. Authors suggested that it could be assigned to the synergistic effect between the retained alkoxyls and the nitrogen doping. The mild one-step approach was adopted to prepare N-doping carbon–TiO2 nanohybrids [58]. Before calcination, a morphism crystal phase was obtained. After calcination, the crystalline structure of the anatase phase and the characteristic peak for C at 26.2° were observed in an XRD pattern (Figure 5a). The XPS spectrum exhibits C1s, O1s, N1s, Ti2p, and Cu2p3/2 (substrate) peaks after calcination (Figure 5b). One can observe the formed N–TiO2 with an average diameter of eight nm and the nicely distributed of the N–carbon dots on N–TiO2 via TEM (Figure 5c), and the junction between N–carbon and N–TiO2 (Figure 5d). The N-doping carbon–TiO2 had a large BET specific surface area of 279.43 m2 g−1, a pore volume of 0.5750 cm3/g, and a pore area of 96.66 m2/g (Figure 6a). N-doping carbon–TiO2 nanohybrids displayed an absorption in the visible band, while P25 showed the band gap absorption onset at 380 nm (Figure 6b). N-doping carbon–TiO2 nanohybrids also had high re/oxidative current peaks, compared with that of P25 (Figure 6c). These N-doping carbon–TiO2 nanohybrids showed excellent photocatalytic ability towards organic (RhB) and inorganic pollutant (K2Cr2O7) degradations under visible light irradiation (Figure 6d).
Using polystyrene spheres as templates and polyaniline as both C and N doping sources, mesoporous C and N co-doped TiO2 hybrid shells were synthesized [59], which exhibited enhanced photocatalytic activity in the degradation of organic dyes and H2 production under visible light irradiation, compared to their undoped counterpart. The co-doped C and N and the presence of graphite carbon were reported to be attributed to the enhanced visible light absorption and charge separation efficiency. They authors observed the synergistic effect in photoelectrochemical reactions exemplified by the concurrent oxidization of perfluorooctanoic acid and hydrogen generation over C–N co-doped TiO2 nanotube arrays as a photoanode [60].
By a thermal treatment of titanium substrates embedded in aluminosilicates or alumina particles, C–N co-doped TiO2 thin film photoanodes were prepared at 500 °C for three hours [61]. The as-prepared C–N co-doped TiO2 photoanodes showed a significant water oxidation with the high photo currents of about 100 mu A cm−2 at zero bias when using alumina as the reactant medium under a standard one SUN illumination of a solar simulator. This photoanode showed a stable photocurrent for about 10 h in the long-term stability test, strong absorption to the visible light region, and improved electron donor density. An in situ carbonization method was also reported to prepare worm-like C–N co-doped TiO2 nanocrystallites with high surface areas for efficient photocatalysis using tetrakis (dimethylamino) titanium as both titanium oxide and nitrogen-dopant sources, and the mesoporous carbon molecular sieves CMK-3 as a hard template as well as carbon doping source [62]. The carbon and nitrogen elements were proven to be responsible for the enhanced visible light-driven photocatalytic activity.

2.3. Photocatalysts of Metal–C Doped TiO2 Composites

Metal–C doped TiO2 composites were reported to improve their photocatalytic activity due to the coupling effect of the incorporation of carbon species and metal nanoparticles. The eggshell membrane templating method was developed to synthesize porous organic carbon-doped titania (C–TiO2) nanomaterials and their composites with Ag nanoparticles (Ag/C–TiO2) [63]. The doped organic carbon was composed of the active carbon and carbonate species, forming a layer around the surface of TiO2 nanoparticles. In the degradation of methylene blue under visible light irradiation, both C–TiO2 and Ag/C–TiO2 nanomaterials showed higher photocatalytic activity compared with the pure TiO2 material, commercial Degussa P25. These results can be assigned to the coupling effect of the incorporation of carbon species and Ag nanoparticles. Ag/C–TiO2 nanomaterials were also prepared by a one-step synthetic approach through the thermal oxidation of an AgNO3-impregnated TiC precursor [64]. The loaded silver nanoparticles were uniformly distributed on the TiO2 surface in the Ag0 metallic state, further enhancing the visible light absorption due to the surface plasmon resonance, yet inhibiting the charge carrier recombination by conduction band electron trapping. The authors proposed the band structure and a tentative visible light photocatalytic disinfection mechanism of Ag/C–TiO2 with four pathways. They demonstrated the enhanced visible light photocatalytic performance for the disinfection of E. coli and E. fecalis in comparison with C–TiO2.
Mo/C co-doped TiO2 powders were reported by thermal oxidizing a mixture of TiC and MoO3 powders in the air [65]. A 32-nm red shift of the spectrum in the visible absorption onset for these materials was observed. Mo/C co-doped TiO2 displayed enhanced photocatalytic activity in the decomposition of methylene blue, and the photocurrent was about four times as high as that of pure anatase TiO2. A calcination-assisted solvothermal reaction was used to prepare the Nd and C co-doped TiO2 (Nd–C–TiO2) without the addition of any other C precursors [66]. The as-prepared sample had a dramatically improved visible absorption, covering the UV and whole visible light range from 200 to 900 nm. Nd–C–TiO2 was reported to represent the excellent destruction ability of NOx gas and degradation activity of methyl orange solution under visible light irradiation, which was superior to pure TiO2, C–TiO2, and Nd–TiO2.
A hydrothermal approach was reported to synthesize C–Fe co-doped TiO2 sheets with dominant {001} facets using TiC and Fe2O3 as the reactants in an HNO3–HF mixed solution [67], resulting in the aggregates of highly truncated bipyramidal microsheets with the length of one to two μm and the thickness of 100–200 nm (Figure 7). The incorporation of Fe into anatase TiO2 did not affect the growth of TiO2 microsheets. The corresponding SAED pattern demonstrated the flat and square surface belonging to the (001) plane of anatase TiO2 (Figure 7a). One can find the (020) and (200) atomic planes with a lattice spacing of 0.19 nm and a vertical angle (Figure 7b). These composites exhibited the narrow band gap, strong visible light absorption, and high separation efficiency of photogenerated carriers, compared with the C-doped TiO2 sheets. The authors suggested that these materials might become promising visible light photocatalysts for organic pollutant degradation and efficient H2 evolution.
The synergistic effect of C–Ag co-doped TiO2 photocatalysts was observed within the GGA plus U framework [68]. C and Ag co-doping could induce a synergistic effect, compared with single C and single Ag doping. After co-doping, the visible absorption was stronger for both the near and far distance co-doped systems than those of the single-doped systems. First-principle calculation was applied to estimate all the band structures, density of states, and absorption spectra of pure, Nd-doped, C-doped, and Nd–C co-doped TiO2 based on the density functional theory [69]. Recently, Ag and Cu co-doped TiO2 deposited on polyurethane was synthesized for the conversion of TiO2 into solar fuels under visible light irradiation [70]. The as-synthesized materials exhibited very high photocatalytic activity for the reduction of gaseous CO2 to produce CH4 and CO fuels. The Ag and Cu dopants also enhanced the separation of the electron–hole pairs of the doped TiO2 photocatalysts, exhibiting the highest photocatalytic reduction of CO2. A silver/carbon co-doped titanium dioxide photocatalyst was also fabricated for improving dye degradation under visible light irradiation through the hydrolysis of titanium tetrachloride followed by calcination at 500 °C [71]. The co-doping of TiO2 with Ag and carbon resulted in an increase in the surface area of the photocatalyst and altered the ratio of anatase to the rutile phase. In comparison with undoped TiO2, doping TiO2 using carbon showed higher photocatalytic activity toward methyl orange degradation under visible light irradiation.

2.4. Photocatalysts of Other Co-Doped C/TiO2 Composites

The synergetic effect is one of the key factors improving the photocatalytic efficiency of co-doped C/TiO2 composites. Xu and Zhang demonstrated the controlled synthesis of C–Cl co-doped anatase TiO2 nanocrystals with mesoporous structures and high surface areas between 158.1–469.6 m2 g−1 by the hydrolysis of titanium tetraisopropoxide in chloroform containing trace amounts of water at a low temperature of 150 °C without a subsequent thermal procedure [72]. C–Cl co-doped anatase TiO2 exhibited higher photoactivity in the photodegradation of RhB under visible light irradiation, compared with C–TiO2 and Cl–TiO2, which was attributed to the larger surface area and the synergetic effect of the co-doping. Hierarchically meso/nanoporous S and C co-doping TiO2 was fabricated by a hydrolysis and calcination method using Ti(OC4H9)4 and K2S2O8 as precursors [73]. The S and C co-doped TiO2 was found to exhibit high photocatalytic H2 generation efficiency in a water/methanol sacrificial reagent system under the irradiation of UV light. Experimental results indicated that the high photocatalytic efficiency is dependent on the comprehensively competing effects of the co-doping of S and C, crystallization, specific surface area, and light absorption capability.
The synthesis of S and C co-doped hierarchically meso/nanoporous TiO2 with excellent thermal stability and photocatalytic activity was reported by Xu et al. [74]. The sulfur nanoparticles in the sulfur hydrosol were applied to serve as the seeds for the formation of bimodal meso/nanopores TiO2 upon calcination treatment. The authors observed remarkably enhanced photocatalytic activity and excellent cyclic stability on the decomposition of methylene blue in the visible light region, which was attributed to the numerous oxygen vacancies, acidic sites on the surface of TiO2, and large specific surface area. The sol–gel method combining with high-energy ball milling was employed to synthesize the (N, S, and C) co-doped TiO2 calcined at the different temperatures (400–700 °C), employing butyl titanate as the titanium source and thiourea as the doping agent [75]. The (N, S, and C) co-doping and the calcination temperature played an important role in the microstructure and photocatalytic activity of the samples, judging from the photocatalytic reduction of Cr(VI) as a model compound under visible light irradiation. As for the control of TiO2-500, it did not cause the concentration decrease of Cr(VI) in aqueous solution upon visible light irradiation for 50 minutes (Figure 8a). In the presence of (N, S, and C) co-doped TiO2-x samples, a quick concentration decrease of Cr(VI) was observed. Moreover, the concentration of Cr(VI) decreased quickly with the increase of calcination temperature from 500 °C to 700 °C. It was found that the decrease in the visible absorption, the separation rate of the electron–hole pairs, the specific surface area, and the increase in crystal size for (N, S, and C) co-doped TiO2-x samples resulted in the decrease in photocatalytic performance (Figure 8b).

3. Photocatalysts of C/g-C3N4 Composites

Graphitic carbon nitride (g-C3N4) is a research hotspot material as a promising photocatalyst due to its good photocatalytic performance and unique electronic structure. A sonochemical method was applied to fabricate the mesoporous g-C3N4/graphene and mesoporous g-C3N4/ graphene oxide nanocomposites with high visible light photocatalytic activity (NO removal ratios of 64.9% and 60.7%, respectively) [76]. The authors suggested that this photocatalytic performance could be attributed to the high surface area and pore volume, improved visible light utilization, enhanced reduction power of electrons, and promoted separation of charge carriers. In Shi’s work, a thermal treatment was also developed during the preparation of the ordered mesoporous carbon/g-C3N4 composites with efficient photocatalytic activity under visible light irradiation [77], which exhibited very high photocatalytic activities in degrading RhB dye, and the high degradation rate was almost 10 times that of pristine g-C3N4, which was partly due to the ordered mesoporous carbon. After five cycles, these composites retained good stability and photodegradation efficiency. The authors suggested that the combined effects came from the enhanced visible light absorption, enriched the dye adsorption on the catalyst, and subsequent efficient separation of the photogenerated electrons and holes. Recently, an in situ synthesis of C-doped TiO2@g-C3N4 core–shell hollow nanospheres with remarkably enhanced visible light photocatalytic activity for water splitting to produce H2 were reported [78], exhibiting the H2 generation rate of 35.6 mu mol g−1 h−1. The heterojunction, in situ growth, and unique hollow structure favored the enhanced photocatalytic performance.
Hydrothermal treatment was also applied to synthesize the C/g-C3N4 composites with a mixture of g-C3N4 and glucose solution [79]. The authors discovered that carbon doping by glucose is an efficient way to improve the photocatalytic performance of g-C3N4 due to the synergistic effect between carbon and g-C3N4, leading to increased visible light absorption, enhanced absorptivity, and improved photoinduced electron–hole separation efficiency. A 97% degradation of Methylene blue (MB) by C2/g-C3N4 composite was observed within 2.0 hours under visible light irradiation, while the pure g-C3N4 only caused a 73% degradation of MB (Figure 9). The photocatalytic activity of C/g-C3N4 composites increased with the weight of the glucose precursor from 0.0025 g to 0.005 g. By adding a small amount of citric acid into urea as the precursor, N-doped graphitic carbon-incorporated g-C3N4 could be synthesized during thermal polymerization [80]. The as-obtained materials retained the original framework of g-C3N4, and showed remarkably enhanced visible light harvesting, and promoted photoexcited charge carrier separation and transfer. They demonstrated an H2 evolution rate of 64 mu mol h-1 on N-doped graphitic carbon-incorporated g-C3N4 under visible light irradiation. This polymerized method was also employed to obtain the visible light-driven N-doped carbon dots/g-C3N4 photocatalyst composite with enhanced visible light photocatalytic activity for the degradation of indomethacin [81]. This enhanced photocatalytic activity might be ascribed to the unique up-converted Photoluminescence (PL) properties, efficient charge separation, as well as band gap narrowing of the N-doped carbon dots. The authors indicated that this composite possesses a very desirable performance for the toxicity reduction and mineralization of indomethacin under long exposures of visible light irradiation.
The precipitation method was introduced to fabricate robust g-C3N4 nanosheets/carbons/NiS multiple heterojunctions for photocatalytic H2 generation [82]. In this work, the authors used the nanocarbons as effective H2-evolution co-catalysts and conductive electron bridges to collect photogenerated electrons, resulting in the downshift of the valence band of g-C3N4 and facilitating the fast oxidation of triethanolamine and charge carrier separation. They also discovered that the amorphous carbons with high electrical conductivity and weak electrocatalytic H2-evolution activity are suitable interfacial bridges between g-C3N4 and NiS co-catalysts for maximizing H2 generation. More recently, a thermal treatment method was employed to synthesize a metal organic framework-derived, nitrogen-doped, carbon-modified g-C3N4 heterostructure composite with enhanced photocatalytic activity for bisphenol A degradation with peroxymonosulfate under visible light irradiation [83], which had the hierarchical porosity, conductive network, and abundant exposed active sites for peroxymonosulfate activation. In this research, the as-prepared heterostructure composite exhibited remarkably improved photocatalytic activities toward bisphenol A degradation in the presence of peroxymonosulfate under visible light irradiation, displaying approximately 8.6 times the apparent rate constant of bare g-C3N4. In general, all these examples mentioned above demonstrated that the carbon doping is an efficient strategy to improve the photocatalytic property of g-C3N4 and that a synergistic effect exists between carbon and g-C3N4.

4. Photocatalysts of Ternary C-Doped Composites

Similar to metal C-doped TiO2 composites, ternary C-doped composites also displayed the synergistic effect from the ternary interfaces. Shen et al. synthesized the magnetic Fe3O4@C @Cu2O composites with a bean-like core/shell nanostructure by a self-assembly approach [84]. They obtained as-prepared magnetic composites with a size of about 420 nm and shells that were composed of several nanoparticles in the range of five to 10 nm. The carbonaceous layer with bound hydrophilic groups was reported to inherit from the starting materials, acting as both the linker and the stabilizer between Fe3O4 and Cu2O. The Fe3O4@C@Cu2O composites were found to exhibit ferromagnetic behavior, good dispersibility in aqueous solution, and universal and powerful visible light photocatalytic activity for the degradation of RhB, methyl orange (MO), and alizarin red (AR), which was superior to commercial Cu2O and Degussa P-25 powders. They suggested that these novel magnetic composites might have applications in dye solution treatment, the degradation of organic pollutants, and environmental cleaning.
The multi-step process of hydrothermal reaction and microwave irradiation was used to synthesize one-dimensional coaxial nanostructures composed of ZnO nanorod cores, intermediate amorphous carbonaceous layers, and CdS nanoparticle sheaths (i.e. ZnO/C/CdS nanocables) [85]. A single coaxial ZnO/C/CdS nanocable was obtained, in which ZnO nanoparticles were encapsulated by an amorphous carbonaceous layer with the thickness of about five nm, and CdS nanoparticles with the size of about 5.5 nm were uniformly deposited on the surface of the carbon layer. In addition, the thickness of the amorphous carbonaceous layer was reported to be easily tuned from three to 10 nm by controlling the amount of glucose from 0.1 to 0.3 g. As for the high concentration of Cd2+, more CdS nanoparticles were anchored onto the amorphous carbonaceous layer, and the size of the CdS nanoparticles became larger with the increasing of the Cd2+ concentration. The thickness of the carbonaceous layer and the amount of CdS nanoparticles can be tuned by controlling the glucose intake and the cadmium salt concentration, respectively. The as-prepared ZnO/C/CdS nanocables exhibited much higher activity than that of pure ZnO nanorods, pure CdS, and ZnO/C core–shell nanocables for the degradation of RhB and methylene blue (MB) under visible light irradiation, which was due to stronger adsorption and the synergistic effect from the ternary ZnO/C/CdS interface.
In the previous literature, Li et al. also found the synergistic effect in sandwiched TiO2@Au@C hollow spheres. The hydrothermal reaction and calcination were applied to the fabrication of uniform double-shelled and sandwiched TiO2@Au@C hollow and porous spheres [86]. An HR-TEM image of the nanocomposite showed the (101) crystallographic planes of TiO2, the (111) plane of Au nanoparticles, and the coated carbon layer with a thickness of ca. 10 nm in Figure 10b,c displayed the presence of individual nanoparticles on a thick and layered shell. It was found that these microspheres displayed high photocatalytic activity and photocatalytic stability for the degradation of 4-nitroaniline, and the remarkable photocatalytic activity for hydrogen generation from methanol/water solutions. TiO2@Au@C achieved a visible light photodegradation of 92.65% for 4-nitroaniline, while P25 only displayed a visible light photodegradation of 17.61% of 4-nitroaniline (Figure 11). It was also observed that the photodegradation reactions followed a pseudo-first-order kinetics with the rate constants (0.0012 min−1, 0.0004 min−1, 0.0050 min−1, 0.0046 min−1, and 0.015 min−1) for 4-nitroaniline and the rate constants (0.0009 min−1, 0.0014 min−1, 0.0036 min−1, 0.0033 min−1, and 0.081 min−1) for 4-nitrophenol by P25, hollow TiO2, hollow TiO2@C, hollow TiO2/Au, and double-shelled and sandwiched TiO2@Au@C hollow spheres, respectively, under visible light irradiation. They proposed the synergistic effect of coupling TiO2 hollow spheres with Au nanoparticles and a C shell in photocatalytic performance.

5. Photocatalysts of Other C-doped Composites

The semiconductor ZnO has been attracting considerable attention because of its large exciton binding energy (60 meV) and possible applications to UV light emitters, gas sensors, transparent electronics, photocatalysis, solar cell, environmental remediation, and surface acoustic wave devices. Novel ZnO@C hybrid composites were obtained using a metal organic framework as a precursor [87]. The resultant ZnO hierarchical aggregates with a three-dimensional cubic morphology and ZnO@C hybrid composites within a highly porous carbonaceous species were obtained. The hybrid composites showed relatively high photocatalytic decomposition activity and significantly enhanced adsorption capacities for organic pollutants. The hydrothermal method was reported for the synthesis of ZnO@C coaxial gemel hexagonal microrods with a thin carbon layer on their surface by using furfural as the carbon precursor, and they showed high photocatalytic activity and stability [88]. The thickness of the surface carbon layer was found to increase with the hydrothermal time until 16 h. The carbon layer was reported to play a significantly important role in the photocatalytic activities and photostabilities of the ZnO@C microrods for the UV-irradiated photodegradation of methylene blue solution. Zhou et al. described the deposition of carbon on the surface of ZnO nanoparticles via an adsorption and calcination process [89]. The as-prepared ZnO/C exhibited much higher photocatalytic properties and adsorption capacity for MB, compared with pure ZnO and carbon, because of the formation of heteroarchitectures, which was tested by photocatalytic decomposition and adsorption of the dye methylene blue in aqueous solution. ZnO/C showed a pseudo-first-order rate constant of 0.326 min−1, which was 10 times larger than that of pure ZnO.
By combining the electrospinning technique and thermal decomposition process, C-doped ZnO nanofibers with high visible light photocatalytic activity were fabricated using the precursor solution consisting of polyacrylonitrile and zinc acetate [90]. The C-doped ZnO nanofibers possessed a much higher photocatalytic degradation rate of RhB compared with the pure ZnO nanofibers under visible light irradiation, which was attributed to the formation of the new energy states in ZnO because of the carbon doping, which might reduce the band gap of the ZnO. By the solvothermal treatment of zinc acetate in an ethylene glycol–ethanol mixture, C-doped ZnO ball-in-ball hollow microspheres were fabricated for photocatalytic and photoelectrochemical applications [91]. The C-doped ZnO ball-in-ball hollow microspheres displayed higher photocatalytic active than ZnO particles by the photodegradation of phenol and MO (Figure 12a). The hollow microspheres exhibited ca. 8.9-fold and 10.5-fold higher activity in comparison with pristine ZnO nanoparticles as the photocatalyst for organic pollutant degradation and as a photoanode material for photoelectrochemical water splitting, respectively (Figure 12a). The C-doped ZnO hollow microspheres also showed stable performance in recycling reactions (Figure 12b). As for the photocurrent–voltage (I–V) curves, C-doped ZnO displayed an obviously high photocurrent (Figure 12c), and an about 10.5 times higher transient photocurrent than ZnO nanoparticles (Figure 12d).
The two-step hydrothermal method was reported to prepare Bi2WO6-coated carbon microspheres using glucose as a carbon source [92]. By the photocatalytic degradation of gas-phase benzene under visible light irradiation, it was revealed that the photocatalytic performance of the Bi2WO6 was greatly enhanced by the introduction of carbon microspheres. The Bi2WO6@C microspheres displayed a photocatalytic benzene conversion of 42.6% and benzene mineralization of 80.0%, respectively. As a direct wide band-gap semiconductor, ZnS nanostructure displayed unique properties, including excellent electrical transport properties, good thermal stability, and high electronic mobility. The density functional theory calculations were used to investigate the origin of the experimentally observed changes in the visible light responsive photoactivity of hexagonal wurtzite ZnS induced by (N, C) co-doping [93]. They performed the accurate comparative analyses of geometric and electronic structures for the different doping models. For the mono N-doped or C-doped ZnS systems, the substituted doping was found to just induce a slight band gap narrowing (less than 0.2 eV), while the interstitial doping led to a more significant optical absorption red shift. For the N–C co-doped systems, the electron transition from the impurity states in the band gap to the conduction band will induce a large red shift, suggesting visible light absorption. The calculated results of 2N–C co-doped ZnS indicated that the trimer doping induces a larger red shift of photoelectron transition and passivates the partially occupied states by the charge compensation effect in a donor–acceptor pair. This work provided a solid basis for the rationalization of the experimentally observed red shift of optical absorption in wurtzite ZnS as a consequence of (N, C) co-doping. Recently, a hydrothermal method was developed to synthesize well-dispersed carbon-doped ZnSn(OH)6 submicrocube photocatalysts at 433 K using glucose as the carbon-doping source [94], which exhibited higher photocatalytic performance by the decomposition of methylene blue in aqueous solution under visible light irradiation, compared with pure ZnSn(OH)6.

6. Conclusions and Future Perspectives

In this review, we summarized the recent advances on the synthesis, photocatalytic properties, and potential applications of photocatalysts based on carbon-based (nano)composites. According to the results from represented examples, these carbon-based (nano)composite photocatalysts exhibited improved photocatalytic properties compared to their individual components. It is worth noting that the synergistic effect is one of the key factors responsible for their photocatalytic performance. Based on our knowledge, some future research trends and directions can be expected as follows. (1) The exploitation of the precursor sources is important for the properties and applications of carbon-based (nano)composites. Various raw materials should be developed as carbon sources. Obviously, different types of biomass such as wood, grass, and bamboo are promising precursor sources of carbon, to which researchers should pay more attention. (2) There have been many examples indicating that the as-prepared carbon-based (nano)composites exhibited improved photocatalytic activity compared with P25 and their individual components. Then, we need to focus on the practical applications of these carbon-based (nano)composites prepared in labs. From an application point of view, it is not enough to obtain carbon-based (nano)composite photocatalysts with only high photocatalytic activity. We should also consider other factors such as the raw materials, production cost, production cycle, yield, and so on. Photocatalytic activity is a key and important factor, but it is not the only factor. It is also important to develop environmentally friendly methods for the large-scale and industrial-scale production of carbon-based nanocomposite photocatalysts at low cost and with high efficiency. (3) There are many examples mentioning the synergistic effect of the incorporation of carbon species in the carbon-based nanocomposite. This effect still needs to be scrutinized in depth, e.g., by using theoretical modeling and computer simulation. Moreover, the relationships between the preparation, structures, properties, and applications of carbon-based nanocomposites need to be dissected in detail. We believe that more and more rapid progresses will be made in the research of carbon-based nanocomposite photocatalysts, since they have broad potential applications in the environmental treatment and renewable energy fields.

Acknowledgments

This research was funded by the Fundamental Research Funds for the Central Universities grant number [No. 2017ZY49], Swedish Energy Agency, Swedish Research Council, and Stiftelsen Olle Engkvist Byggmästare.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, J.L.; Wu, Y.M.; Xing, M.Y.; Leghari, S.A.K.; Sajjad, S. Development of modified N doped TiO2 photocatalyst with metals, nonmetals and metal oxides. Energy Environ. Sci. 2010, 3, 715–726. [Google Scholar] [CrossRef]
  3. Zhu, J.F.; Zäch, M. Nanostructured materials for photocatalytic hydrogen production. Curr. Opin. Coll. Interface Sci. 2009, 14, 260–269. [Google Scholar] [CrossRef]
  4. Zhu, J.F.; Chakarov, D.; Zäch, M. Chapter 13: Nanostructured Materials for Photolytic Hydrogen Production. In Energy Efficiency and Renewable Energy through Nanotechnology; Zang, L., Ed.; Springer: London, UK, 2011; pp. 441–486. ISBN 978-0-85729-637-5. [Google Scholar]
  5. Park, H.; Park, Y.; Kim, W.; Choi, W. Surface modification of TiO2 photocatalyst for environmental applications. J. Photochem. Photobiol. C-Photochem. Rev. 2013, 15, 1–20. [Google Scholar] [CrossRef]
  6. Lee, K.M.; Lai, C.W.; Ngai, K.S.; Juan, J.C. Recent developments of zinc oxide based photocatalyst in water treatment technology: A review. Water Res. 2016, 88, 428–448. [Google Scholar] [CrossRef] [PubMed]
  7. Zhu, J.F. Photocatalytic hydrogen production. In Encyclopedia of Sustainability Science and Technology; Meyers, R.A., Ed.; Springer: New York, NY, USA, 2012; pp. 7881–7901. ISBN 978-0-387-89469-0. [Google Scholar]
  8. Zhu, J.F. Photocatalysts for hydrogen production. In Advanced Materials for Clean Energy; Xu, Q., Kobayashi, T., Eds.; CRC Press, Taylor & Francis Group: Boca Raton, FL, USA, 2015; pp. 391–420. ISBN 978-1-4822-0578-7. [Google Scholar]
  9. Kumar, S.G.; Devi, L.G. Review on modified TiO2 photocatalysis under UV/visible light: Selected results and related mechanisms on interfacial charge carrier transfer dynamics. J. Phys. Chem. A 2011, 115, 13211–13241. [Google Scholar] [CrossRef] [PubMed]
  10. Kim, J.; Lee, C.W.; Choi, W. Platinized WO3 as an environmental photocatalyst that generates oh radicals under visible light. Environ. Sci. Technol. 2010, 44, 6849–6854. [Google Scholar] [CrossRef] [PubMed]
  11. Gu, X.Q.; Li, C.Y.; Yuan, S.; Ma, M.G.; Qiang, Y.H.; Zhu, J.F. ZnO based heterojunctions and their application in environmental photocatalysis. Nanotechnology 2016, 27, 402001. [Google Scholar] [CrossRef]
  12. Zhu, C.; Liu, C.G.; Zhou, Y.J.; Fu, Y.J.; Guo, S.J.; Li, H.; Zhao, S.Q.; Huang, H.; Liu, Y.; Kang, Z.H. Carbon dots enhance the stability of CdS for visible-light-driven overall water splitting. Appl. Catal. B-Environ. 2017, 216, 114–121. [Google Scholar] [CrossRef]
  13. Reddy, D.A.; Park, H.; Ma, R.; Kumar, D.P.; Lim, M.; Kim, T.K. Heterostructured WS2-MoS2 ultrathin nanosheets integrated on CdS nanorods to promote charge separation and migration and improve solar-driven photocatalytic hydrogen evolution. Chemsuschem 2017, 10, 1563–1570. [Google Scholar] [CrossRef]
  14. Li, Y.; Wang, L.L.; Cai, T.; Zhang, S.Q.; Liu, Y.T.; Song, Y.Z.; Dong, X.R.; Hu, L. Glucose-assisted synthesize 1D/2D nearly vertical CdS/MoS2 heterostructures for efficient photocatalytic hydrogen evolution. Chem. Eng. J. 2017, 321, 366–374. [Google Scholar] [CrossRef]
  15. Cao, S.W.; Low, J.X.; Yu, J.G.; Jaroniec, M. Polymeric photocatalysts based on graphitic carbon nitride. Adv. Mater. 2015, 27, 2150–2176. [Google Scholar] [CrossRef]
  16. Yan, S.C.; Lv, S.B.; Li, Z.S.; Zou, Z.G. Organic-inorganic composite photocatalyst of g-C3N4 and TaON with improved visible light photocatalytic activities. Dalton Trans. 2010, 39, 1488–1491. [Google Scholar] [CrossRef]
  17. Pan, C.S.; Xu, J.; Wang, Y.J.; Li, D.; Zhu, Y.F. Dramatic activity of C3N4/BiPO4 photocatalyst with core/shell structure formed by self-assembly. Adv. Funct. Mater. 2012, 22, 1518–1524. [Google Scholar] [CrossRef]
  18. Li, T.B.; Chen, G.; Zhou, C.; Shen, Z.Y.; Jin, R.C.; Sun, J.X. New photocatalyst BiOCl/BiOI composites with highly enhanced visible light photocatalytic performances. Dalton Trans. 2011, 40, 6751–6758. [Google Scholar] [CrossRef]
  19. Liu, R.H.; Huang, H.; Li, H.T.; Liu, Y.; Zhong, J.; Li, Y.Y.; Zhang, S.; Kang, Z.H. Metal nanoparticle/carbon quantum dot composite as a photocatalyst for high-efficiency cyclohexane oxidation. ACS Catal. 2014, 4, 328–336. [Google Scholar] [CrossRef]
  20. Upadhyay, R.K.; Soin, N.; Roy, S.S. Role of graphene/metal oxide composites as photocatalysts, adsorbents and disinfectants in water treatment: A review. RSC Adv. 2014, 4, 3823–3851. [Google Scholar] [CrossRef]
  21. Xiang, Q.J.; Yu, J.G.; Jaroniec, M. Graphene-based semiconductor photocatalysts. Chem. Soc. Rev. 2012, 41, 782–796. [Google Scholar] [CrossRef]
  22. Zhang, N.; Liu, S.Q.; Xu, Y.J. Recent progress on metal core@semiconductor shell nanocomposites as a promising type of photocatalyst. Nanoscale 2012, 4, 2227–2238. [Google Scholar] [CrossRef]
  23. Zhang, H.; Lv, X.J.; Li, Y.M.; Wang, Y.; Li, J.H. P25-graphene composite as a high performance photocatalyst. ACS Nano 2010, 4, 380–386. [Google Scholar] [CrossRef]
  24. Giovannetti, R.; Rommozzi, E.; Zannotti, M.; D’Amato, C.A.; Ferraro, S.; Cespi, M.; Bonacucina, G.; Minicucci, M.; Di Cicco, A. Exfoliation of graphite into graphene in aqueous solution: an application as graphene/TiO2 nanocomposite to improve visible light photocatalytic activity. RSC Adv. 2016, 6, 93048–93055. [Google Scholar] [CrossRef]
  25. Giovannetti, R.; Rommozzi, E.; Zannotti, M.; D’Amato, C.A. Recent advances in graphene based TiO2 nanocomposites (GTiO2Ns) for photocatalytic degradation of synthetic dyes. Catalysts 2017, 7, 305. [Google Scholar] [CrossRef]
  26. Rommozzi, E.; Zannotti, M.; Giovannetti, R.; D’Amato, C.A.; Ferraro, S.; Minicucci, M.; Gunnella, R.; Di Cicco, A. Reduced graphene oxide/TiO2 nanocomposite: from synthesis to characterization for efficient visible light photocatalytic applications. Catalysts 2018, 8, 598. [Google Scholar] [CrossRef]
  27. Zhang, N.; Zhang, Y.H.; Xu, Y.J. Recent progress on graphene-based photocatalysts: Current status and future perspectives. Nanoscale 2012, 4, 5792–5813. [Google Scholar] [CrossRef] [PubMed]
  28. Akhavan, O. Graphene nanomesh by ZnO nanorod photocatalysts. ACS Nano 2010, 4, 4174–4180. [Google Scholar] [CrossRef] [PubMed]
  29. Zhang, N.; Yang, M.Q.; Liu, S.Q.; Sun, Y.G.; Xu, Y.J. Waltzing with the versatile platform of graphene to synthesize composite photocatalysts. Chem. Rev. 2015, 115, 10307–10377. [Google Scholar] [CrossRef] [PubMed]
  30. Shanmugam, S.; Gabashvili, A.; Jacob, D.S.; Yu, J.C.; Gedanken, A. Synthesis and characterization of TiO2@C core-shell composite nanoparticles and evaluation of their photocatalytic activities. Chem. Mater. 2006, 18, 2275–2282. [Google Scholar] [CrossRef]
  31. Sullivan, J.A.; Neville, E.M.; Herron, R.; Thampi, K.R.; Donal MacElroy, J.M. Routes to visible light active C-doped TiO2 photocatalysts using carbon atoms from the Ti precursors. J. Photochem. Photobiol. A Chem. 2014, 289, 60–65. [Google Scholar] [CrossRef]
  32. Moghaddam, H.A.; Jafari, S.; Mohammadi, M.R. Enhanced efficiency of over 10% in dye-sensitized solar cells through C and N single- and co-doped TiO2 single-layer electrodes. New J. Chem. 2017, 41, 9453–9460. [Google Scholar] [CrossRef]
  33. Noorimotlagh, Z.; Kazeminezhad, I.; Jaafarzadeh, N.; Ahmadi, M.; Ramezani, Z.; Martinez, S.S. The visible-light photodegradation of nonylphenol in the presence of carbon-doped TiO2 with rutile/anatase ratio coated on GAC: Effect of parameters and degradation mechanism. J. Hazard. Mater. 2018, 350, 108–120. [Google Scholar] [CrossRef]
  34. Wu, X.Y.; Yin, S.; Dong, Q.; Sato, T. Blue/green/red colour emitting up-conversion phosphors coupled C-TiO2 composites with UV, visible and NIR responsive photocatalytic performance. Appl. Catal. B Environ. 2014, 156–157, 257–264. [Google Scholar] [CrossRef]
  35. Huang, Q.W.; Tian, S.Q.; Zeng, D.W.; Wang, X.X.; Song, W.L.; Li, Y.Y.; Xiao, W.; Xie, C.S. Enhanced photocatalytic activity of chemically bonded TiO2/graphene composites based on the effective interfacial charge transfer through C-Ti bond. ACS Catal. 2013, 3, 1477–1485. [Google Scholar] [CrossRef]
  36. Wang, S.; Zhao, L.; Bai, L.N.; Yan, J.M.; Jiang, Q.; Lian, J.S. Enhancing photocatalytic activity of disorder engineered C/TiO2 and TiO2 nanoparticles. J. Mater. Chem. A 2014, 2, 7439–7445. [Google Scholar] [CrossRef]
  37. Zhou, W.; Liu, Y.; Zhang, Y.Z.; Yang, G.; Deng, S.H.; Shen, F.; Peng, H.; Wang, L.L. Novel multi-layer cross-linked TiO2/C nanosheets and their photocatalytic properties. New J. Chem. 2014, 38, 1647–1654. [Google Scholar] [CrossRef]
  38. Lin, Y.T.; Weng, C.H.; Chen, F.Y. Key operating parameters affecting photocatalytic activity of visible-light-induced C-doped TiO2 catalyst for ethylene oxidation. Chem. Eng. J. 2014, 248, 175–183. [Google Scholar] [CrossRef]
  39. Hassan, M.E.; Cong, L.; Liu, G.L.; Zhu, D.W.; Cai, J.B. Synthesis and characterization of C-doped TiO2 thin films for visible-light-induced photocatalytic degradation of methyl orange. Appl. Surf. Sci. 2014, 294, 89–94. [Google Scholar] [CrossRef]
  40. Lu, J.; Wang, Y.; Huang, J.F.; Fei, J.; Cao, L.Y.; Li, C.Y. In situ synthesis of mesoporous C-doped TiO2 single crystal with oxygen vacancy and its enhanced sunlight photocatalytic properties. Dyes Pigments 2017, 144, 203–211. [Google Scholar] [CrossRef]
  41. Dong, F.; Wang, H.Q.; Wu, Z.B. One-step “green” synthetic approach for mesoporous c-doped titanium dioxide with efficient visible light photocatalytic activity. J. Phys. Chem. C 2009, 113, 16717–16723. [Google Scholar] [CrossRef]
  42. Dong, F.; Guo, S.; Wang, H.Q.; Li, X.F.; Wu, Z.B. Enhancement of the visible light photocatalytic activity of C-doped TiO2 nanomaterials prepared by a green synthetic approach. J. Phys. Chem. C 2011, 115, 13285–13292. [Google Scholar] [CrossRef]
  43. Zhou, G.B.; Liu, X.W.; Nan, C.Y.; Liu, Y.X.; Wang, D.S.; Chen, X.Q. C/N-sensitized self-assembly of mesostructured TiO2 nanospheres with significantly enhanced photocatalytic activity. New J. Chem. 2013, 37, 2582–2588. [Google Scholar] [CrossRef]
  44. Shi, J.W.; Chen, J.W.; Cui, H.J.; Fu, M.L.; Luo, H.Y.; Xu, B.; Ye, Z.L. One template approach to synthesize C-doped titania hollow spheres with high visible-light photocatalytic activity. Chem. Eng. J. 2012, 195–196, 226–232. [Google Scholar] [CrossRef]
  45. Zhang, Y.; Zhao, Z.Y.; Chen, J.R.; Cheng, L.; Chang, J.; Sheng, W.C.; Hu, C.Y.; Cao, S.S. C-doped hollow TiO2 spheres: In situ synthesis, controlled shell thickness, and superior visible-light photocatalytic activity. Appl. Catal. B Environ. 2015, 165, 715–722. [Google Scholar] [CrossRef]
  46. Xie, C.; Yang, S.H.; Li, B.B.; Wang, H.K.; Shi, J.W.; Li, G.D.; Niu, C.M. C-doped mesoporous anatase TiO2 comprising 10 nm crystallites. J. Coll. Interface Sci. 2016, 476, 1–8. [Google Scholar] [CrossRef] [PubMed]
  47. Purbia, R.; Borah, R.; Paria, S. Carbon-doped mesoporous anatase TiO2 multi-tubes nanostructures for highly improved visible light photocatalytic activity. Inorg. Chem. 2017, 56, 10107–10116. [Google Scholar] [CrossRef] [PubMed]
  48. Shi, Z.J.; Ma, M.G. Synthesis, structure, and applications of lignin-based carbon materials: A review. Sci. Adv. Mater. 2019, 11, 18–32. [Google Scholar] [CrossRef]
  49. Liu, S.; Ma, C.; Ma, M.G.; Li, J.F. Recent advances in carbon nanomaterials derived from biomass. Sci. Adv. Mater. 2019, 11, 5–17. [Google Scholar] [CrossRef]
  50. Liu, S.; Fu, L.H.; Liu, Y.J.; Meng, L.Y.; Dong, Y.Y.; Li, Y.Y.; Ma, M.G. Cu/C or Cu2O/C composites: Selective synthesis, characterization, and applications in water treatment. Sci. Adv. Mater. 2016, 8, 2045–2053. [Google Scholar] [CrossRef]
  51. Wang, B.; Liu, B.; Ji, X.X.; Ma, M.G. Synthesis, characterization, and photocatalytic properties of bamboo charcoal/TiO2 composites using four sizes powder. Materials 2018, 11, 670. [Google Scholar] [CrossRef]
  52. Liu, B.; Liu, S.; Meng, L.Y.; Li, Y.Y.; Wang, B.; Ma, M.G. Microwave-hydrothermal synthesis and photocatalytic properties of biomass charcoal/TiO2 nanocomposites. J. Saudi Chem. Soc. 2018, 22, 509–518. [Google Scholar] [CrossRef]
  53. Liu, X.; Li, Y.L.; Yang, J.; Wang, B.; Ma, M.G.; Xu, F.; Sun, R.C.; Zhang, X.M. Enhanced photocatalytic activity of CdS-decorated TiO2/carbon core-shell microspheres derived from microcrystalline cellulose. Materials 2016, 9, 245. [Google Scholar] [CrossRef]
  54. Wang, X.P.; Lim, T.T. Solvothermal synthesis of C–N codoped TiO2 and photocatalytic evaluation for bisphenol A degradation using a visible-light irradiated LED photoreactor. Appl. Catal. B Environ. 2010, 100, 355–364. [Google Scholar] [CrossRef]
  55. Dolat, D.; Quici, N.; Kusiak-Nejman, E.; Morawski, A.W.; Puma, G.L. One-step, hydrothermal synthesis of nitrogen, carbon co-doped titanium dioxide (N,C TiO2) photocatalysts. Effect of alcohol degree and chain length as carbon dopant precursors on photocatalytic activity and catalyst deactivation. Appl. Catal. B Environ. 2012, 115–116, 81–89. [Google Scholar] [CrossRef]
  56. Wang, D.H.; Jia, L.; Wu, X.L.; Lu, L.Q.; Xu, A.W. One-step hydrothermal synthesis of N-doped TiO2/C nanocomposites with high visible light photocatalytic activity. Nanoscale 2012, 4, 576–584. [Google Scholar] [CrossRef]
  57. Wu, D.Y.; Wang, L.Z. Low-temperature synthesis of anatase C-N-TiO2 photocatalyst with enhanced visible-light-induced photocatalytic activity. Appl. Surf. Sci. 2013, 271, 357–361. [Google Scholar] [CrossRef]
  58. Ming, H.; Huang, H.; Pan, K.M.; Li, H.T.; Liu, Y.; Kang, Z.H. C/TiO2 nanohybrids co-doped by N and their enhanced photocatalytic ability. J. Solid State Chem. 2012, 192, 305–311. [Google Scholar] [CrossRef]
  59. Wang, M.G.; Han, J.; Hu, Y.M.; Guo, R. Mesoporous C, N-codoped TiO2 hybrid shells with enhanced visible light photocatalytic performance. RSC Adv. 2017, 7, 15513–15520. [Google Scholar] [CrossRef]
  60. Peng, Y.P.; Chen, H.L.; Huang, C.P. The synergistic effect of photoelectrochemical (PEC) reactions exemplified by concurrent perfluorooctanoic acid (PFOA) degradation and hydrogen generation over carbon and nitrogen codoped TiO2 nanotube arrays (C-N-TNTAs) photoelectrode. Appl. Catal. B-Environ. 2017, 209, 437–446. [Google Scholar] [CrossRef]
  61. Venkatkarthick, R.; Davidson, D.J.; Vasudevan, S.; Sozhan, G.; Ravichandran, S. An investigation of interfacial and photoelectrochemical performance of thermally prepared C,N-codoped TiO2 photoanodes for water splitting. Chemistryselect 2017, 2, 288–294. [Google Scholar] [CrossRef]
  62. Zhang, X.Y.; Zhang, Y.; Zhou, L.G.; Li, X.K.; Guo, X.Y. In situ C,N-codoped mesoporous TiO2 nanocrystallites with high surface areas and worm-like structure for efficient photocatalysis. J. Porous Mater. 2018, 25, 571–579. [Google Scholar] [CrossRef]
  63. Wang, Q.; Jiang, Z.Y.; Wang, Y.B.; Chen, D.M.; Yang, D. Photocatalytic properties of porous C-doped TiO2 and Ag/C-doped TiO2 nanomaterials by eggshell membrane templating. J. Nanopart. Res. 2009, 11, 375–384. [Google Scholar] [CrossRef]
  64. Zhang, L.; Han, M.D.; Tan, O.K.; Tse, M.S.; Wang, Y.X.; Sze, C.C. Facile fabrication of Ag/C-TiO2 nanoparticles with enhanced visible light photocatalytic activity for disinfection of Escherichia coli and Enterococcus faecalis. J. Mater. Chem. B 2013, 1, 564–570. [Google Scholar] [CrossRef]
  65. Zhang, J.; Pan, C.X.; Fang, P.F.; Wei, J.H.; Xiong, R. Mo + C co-doped TiO2 using thermal oxidation for enhancing photocatalytic activity. ACS Appl. Mater. Interfaces 2010, 2, 1173–1176. [Google Scholar] [CrossRef] [PubMed]
  66. Wu, X.Y.; Yin, S.; Dong, Q.; Guo, C.S.; Kimura, T.; Matsushita, J.; Sato, T. Photocatalytic properties of Nd and C codoped TiO2 with the whole range of visible light absorption. J. Phys. Chem. C 2013, 117, 8345–8352. [Google Scholar] [CrossRef]
  67. Yang, D.; Li, Y.B.; Tong, Z.W.; Sun, Y.Y.; Jiang, Z.Y. One-pot fabrication of C-Fe-co-doped TiO2 sheets with dominant {001} facets for enhanced visible-light photocatalytic activity. Ind. Engineer. Chem. Res. 2014, 53, 19249–19256. [Google Scholar]
  68. Guo, M.L. Synergistic effect of C, Ag-codoped TiO2 photocatalyst within the GGA plus U framework. RSC Adv. 2015, 5, 434–439. [Google Scholar] [CrossRef]
  69. Wu, Y.; Tian, Y.; Zheng, S.K. First principles study on the electronic structure and optical property of Nd-C codoped anatase TiO2. Mater.-Rio De Janeiro 2016, 21, 301–306. [Google Scholar]
  70. Pham, T.D.; Lee, B.K. Novel capture and photocatalytic conversion of CO2 into solar fuels by metals co-doped TiO2 deposited on PU under visible light. Appl. Catal. A-Gen. 2017, 529, 40–48. [Google Scholar] [CrossRef]
  71. Nyamukamba, P.; Tichagwa, L.; Mamphweli, S.; Petrik, L. Silver/carbon codoped titanium dioxide photocatalyst for improved dye degradation under visible light. Int. J. Photoenergy 2017, 2017, 3079276. [Google Scholar] [CrossRef]
  72. Xu, H.; Zhang, L.Z. Controllable one-pot synthesis and enhanced visible light photocatalytic activity of Tunable C-Cl-Co-doped TiO2 nanocrystals with high surface area. J. Phys. Chem. C 2010, 114, 940–946. [Google Scholar] [CrossRef]
  73. Bai, H.W.; Kwan, K.S.Y.; Liu, Z.Y.; Song, X.X.; Lee, S.S.; Sun, D.D. Facile synthesis of hierarchically meso/nanoporous S- and C-co-doped TiO2 and its high photocatalytic efficiency in H2 generation. Appl. Catal. B Environ. 2013, 129, 294–300. [Google Scholar] [CrossRef]
  74. Xu, P.; Xu, T.; Lu, J.; Gao, S.M.; Hosmane, N.S.; Huang, B.B.; Dai, Y.; Wang, Y.B. Visible-light-driven photocatalytic S- and C- codoped meso/nanoporous TiO2. Energy Environ. Sci. 2010, 3, 1128–1134. [Google Scholar] [CrossRef]
  75. Lei, X.F.; Xue, X.X.; Yang, H.; Chen, C.; Li, X.; Niu, M.C.; Gao, X.Y.; Yang, Y.T. Effect of calcination temperature on the structure and visible-light photocatalytic activities of (N, S and C) co-doped TiO2 nano-materials. Appl. Surf. Sci. 2015, 332, 172–180. [Google Scholar] [CrossRef]
  76. Li, Y.H.; Sun, Y.J.; Dong, F.; Ho, W.K. Enhancing the photocatalytic activity of bulk g-C3N4 by introducing mesoporous structure and hybridizing with grapheme. J. Coll. Interface Sci. 2014, 436, 29–36. [Google Scholar] [CrossRef] [PubMed]
  77. Shi, L.; Liang, L.; Ma, J.; Wang, F.X.; Sun, J.M. Remarkably enhanced photocatalytic activity of ordered mesoporous carbon/g-C3N4 composite photocatalysts under visible light. Dalton Trans. 2014, 43, 7236–7244. [Google Scholar] [CrossRef]
  78. Zou, Y.J.; Shi, J.W.; Ma, D.D.; Fan, Z.Y.; Lu, L.; Niu, C.M. In situ synthesis of C-doped TiO2@g-C3N4 core-shell hollow nanospheres with enhanced visible-light photocatalytic activity for H2 evolution. Chem. Eng. J. 2017, 322, 435–444. [Google Scholar] [CrossRef]
  79. Li, Y.P.; Wu, S.L.; Huang, L.Y.; Wang, J.L.; Xu, H.; Li, H.M. Synthesis of carbon-doped g-C3N4 composites with enhanced visible-light photocatalytic activity. Mater. Lett. 2014, 137, 281–284. [Google Scholar] [CrossRef]
  80. 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]
  81. 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. 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]
  82. Wen, J.Q.; Xie, J.; Yang, Z.H.; Shen, R.C.; Li, H.Y.; Luo, X.Y.; Chen, X.B.; Li, X. Fabricating the Robust g-C3N4 Nanosheets/carbons/NiS multiple heterojunctions for enhanced photocatalytic H2 generation: An insight into the trifunctional roles of nanocarbons. ACS Sustain. Chem. Eng. 2017, 5, 2224–2236. [Google Scholar] [CrossRef]
  83. Gong, Y.; Zhao, X.; Zhang, H.; Yang, B.; Xiao, K.; Guo, T.; Zhang, J.J.; Shao, H.X.; Wang, Y.B.; Yu, G. MOF-derived nitrogen doped carbon modified g-C3N4 heterostructure composite with enhanced photocatalytic activity for bisphenol A degradation with peroxymonosulfate under visible light irradiation. Appl. Catal. B-Environ. 2018, 233, 35–45. [Google Scholar] [CrossRef]
  84. Li, S.K.; Huang, F.Z.; Wang, Y.; Shen, Y.H.; Qiu, L.G.; Xie, A.J.; Xu, S.J. Magnetic Fe3O4@C@Cu2O composites with bean-like core/shell nanostructures: Synthesis, properties and application in recyclable photocatalytic degradation of dye pollutants. J. Mater. Chem. 2011, 21, 7459–7466. [Google Scholar] [CrossRef]
  85. Zhou, M.J.; Hu, Y.; Liu, Y.; Yang, W.L.; Qian, H.S. Microwave-assisted route to fabricate coaxial ZnO/C/CdS nanocables with enhanced visible light-driven photocatalytic activity. CrystEngComm 2012, 14, 7686–7693. [Google Scholar] [CrossRef]
  86. Cai, J.B.; Wu, X.Q.; Li, S.X.; Zheng, F.Y.; Zhu, L.C.; Lai, Z.H. Synergistic effect of double-shelled and sandwiched TiO2@Au@C hollow spheres with enhanced visible-light-driven photocatalytic activity. ACS Appl. Mater. Interfaces 2015, 7, 3764–3772. [Google Scholar] [CrossRef]
  87. Yang, S.J.; Im, J.H.; Kim, T.; Lee, K.; Park, C.R. MOF-derived ZnO and ZnO@C composites with high photocatalytic activity and adsorption capacity. J. Hazard. Mater. 2011, 186, 376–382. [Google Scholar] [CrossRef] [PubMed]
  88. Zhang, P.; Li, B.B.; Zhao, Z.B.; Yu, C.; Hu, C.; Wu, S.J.; Qiu, J.S. Furfural-induced hydrothermal synthesis of ZnO@C gemel hexagonal microrods with enhanced photocatalytic activity and stability. ACS Appl. Mater. Interfaces 2014, 6, 8560–8566. [Google Scholar] [CrossRef] [PubMed]
  89. Ma, S.S.; Xue, J.J.; Zhou, Y.M.; Zhang, Z.W.; Wu, X. A facile route for the preparation of ZnO/C composites with high photocatalytic activity and adsorption capacity. CrystEngComm 2014, 16, 4478. [Google Scholar] [CrossRef]
  90. Mu, J.B.; Guo, Z.C.; Che, H.W.; Zhang, X.L.; Bai, Y.M.; Hou, J.X. Electrospinning of C-doped ZnO nanofibers with high visible-light photocatalytic activity. J. Sol-Gel Sci. Technol. 2016, 78, 99–109. [Google Scholar] [CrossRef]
  91. Wang, S.B.; Zhang, X.W.; Li, S.; Fang, Y.; Pan, L.; Zou, J.J. C-doped ZnO ball-in-ball hollow microspheres for efficient photocatalytic and photoelectrochemical applications. J. Hazard. Mater. 2017, 331, 235–245. [Google Scholar] [CrossRef]
  92. Chen, Y.L.; Cao, X.X.; Kuang, J.D.; Chen, Z.; Chen, J.L.; Lin, B.Z. The gas-phase photocatalytic mineralization of benzene over visible-light-driven Bi2WO6@C microspheres. Catal. Commun. 2010, 12, 247–250. [Google Scholar] [CrossRef]
  93. Sun, H.G.; Zhao, X.; Zhang, L.; Fan, W.L. Origin of the enhanced visible photocatalytic activity in (N, C)-codoped ZnS studied from density functional theory. J. Phys. Chem. C 2011, 115, 2218–2227. [Google Scholar] [CrossRef]
  94. Lu, H.X.; Lei, J.; Li, X.X.; Shao, G.; Hou, T.C.; Fan, B.B.; Chen, D.L.; Zhang, L.W.; Wang, H.L.; Xu, H.L. Synthesis and characterization of carbon-doped ZnSn(OH)6 with enhanced photoactivity by hydrothermal method. Cryst. Res. Technol. 2016, 51, 11–15. [Google Scholar] [CrossRef]
Figure 1. (a,b) TEM and (c,d) HR-TEM images of G2.5-TiO2 nanocomposite. From Ref. [35]. Reprinted with permission from American Chemical Society (ACS).
Figure 1. (a,b) TEM and (c,d) HR-TEM images of G2.5-TiO2 nanocomposite. From Ref. [35]. Reprinted with permission from American Chemical Society (ACS).
Catalysts 09 00020 g001
Figure 2. Schematic illustrations for the interfacial charge transfer effect in TiO2/graphene composites: (a) the chemical bonding between TiO2 nanoparticles and graphene nanosheets; (b) the reported energy level of the valence band (VB) and conduction band (CB) vs. NHE (normal hydrogen electrode). From Ref. [35]. Reprinted with permission from ACS.
Figure 2. Schematic illustrations for the interfacial charge transfer effect in TiO2/graphene composites: (a) the chemical bonding between TiO2 nanoparticles and graphene nanosheets; (b) the reported energy level of the valence band (VB) and conduction band (CB) vs. NHE (normal hydrogen electrode). From Ref. [35]. Reprinted with permission from ACS.
Catalysts 09 00020 g002
Figure 3. HR-TEM micrographs of (a) reference TiO2, (b) C/TiO2, and (c) reduced TiO2, along with Fourier-filtered images (a1, b1, and c1) of selected areas and local enlarged images (b2, b3, c2, and c3) shown as inserts. (d) Electron Energy Loss (EELS) spectra of the Ti-L2;3 edge for the three samples. From Ref. [36]. Reprinted with permission from Royal Society of Chemistry (RSC).
Figure 3. HR-TEM micrographs of (a) reference TiO2, (b) C/TiO2, and (c) reduced TiO2, along with Fourier-filtered images (a1, b1, and c1) of selected areas and local enlarged images (b2, b3, c2, and c3) shown as inserts. (d) Electron Energy Loss (EELS) spectra of the Ti-L2;3 edge for the three samples. From Ref. [36]. Reprinted with permission from Royal Society of Chemistry (RSC).
Catalysts 09 00020 g003
Figure 4. SEM and TEM images of the C-doped hollow TiO2 spheres produced with different amounts of tetrabutyl orthotitanate: (a,b) 2 g; (c,d) 3 g; (e,f) 4 g; and (g,h) 5 g. From Ref. [45]. Reprinted with permission from Elsevier.
Figure 4. SEM and TEM images of the C-doped hollow TiO2 spheres produced with different amounts of tetrabutyl orthotitanate: (a,b) 2 g; (c,d) 3 g; (e,f) 4 g; and (g,h) 5 g. From Ref. [45]. Reprinted with permission from Elsevier.
Catalysts 09 00020 g004
Figure 5. (a) XRD patterns and (b) XPS full spectrum of N-doping carbon–TiO2 nanohybrids. (c) TEM image. Inset: digital image of N-doping carbon–TiO2 nanohybrids solution. (d) HR-TEM image. Inset: HR-TEM image of single N–TiO2 on N doped C film (N–Carbon). From Ref. [58]. Reprinted with permission from Elsevier.
Figure 5. (a) XRD patterns and (b) XPS full spectrum of N-doping carbon–TiO2 nanohybrids. (c) TEM image. Inset: digital image of N-doping carbon–TiO2 nanohybrids solution. (d) HR-TEM image. Inset: HR-TEM image of single N–TiO2 on N doped C film (N–Carbon). From Ref. [58]. Reprinted with permission from Elsevier.
Catalysts 09 00020 g005
Figure 6. (a) N2 adsorption–desorption isotherm of N-doping carbon–TiO2; (b) UV-VIS diffuse reflectance absorption spectra of P25, N-doping carbon–TiO2 and physical mix; (c) CV spectra of P25 and N-doping carbon–TiO2; (d) Photodegradation of rhodamine B (RhB) and K2Cr2O7 over P25 and N-doping carbon–TiO2. From Ref. [58]. Reprinted with permission from Elsevier.
Figure 6. (a) N2 adsorption–desorption isotherm of N-doping carbon–TiO2; (b) UV-VIS diffuse reflectance absorption spectra of P25, N-doping carbon–TiO2 and physical mix; (c) CV spectra of P25 and N-doping carbon–TiO2; (d) Photodegradation of rhodamine B (RhB) and K2Cr2O7 over P25 and N-doping carbon–TiO2. From Ref. [58]. Reprinted with permission from Elsevier.
Catalysts 09 00020 g006
Figure 7. TEM image (a), HR-TEM image (b), and SEM image (c,d) of CFTS-1.0. The inset in Figure 6a is the selected area electron diffraction pattern. From Ref. [67]. Reprinted with permission from ACS.
Figure 7. TEM image (a), HR-TEM image (b), and SEM image (c,d) of CFTS-1.0. The inset in Figure 6a is the selected area electron diffraction pattern. From Ref. [67]. Reprinted with permission from ACS.
Catalysts 09 00020 g007
Figure 8. (a) Variation of Cr(VI) concentration (pH 1.5 and C0 = 10 mg/L) with irradiation time in the presence of TiO2–500 and (N, S, and C) co-doped TiO2-x samples under visible light (λ > 420 nm) irradiation, together with the results from the blank experiments (no photocatalyst, but with visible light irradiation); (b) plots of [(C0 − Ct)/C0, t = 50 min] in comparison with Eg values, specific surface area, and crystal size for all (N, S, and C) co-doped-TiO2-x samples. From Ref. [75]. Reprinted with permission from Elsevier.
Figure 8. (a) Variation of Cr(VI) concentration (pH 1.5 and C0 = 10 mg/L) with irradiation time in the presence of TiO2–500 and (N, S, and C) co-doped TiO2-x samples under visible light (λ > 420 nm) irradiation, together with the results from the blank experiments (no photocatalyst, but with visible light irradiation); (b) plots of [(C0 − Ct)/C0, t = 50 min] in comparison with Eg values, specific surface area, and crystal size for all (N, S, and C) co-doped-TiO2-x samples. From Ref. [75]. Reprinted with permission from Elsevier.
Catalysts 09 00020 g008
Figure 9. Photocatalytic degradation efficiency of Methylene blue (MB) by graphitic carbon nitride (g-C3N4) and C/g-C3N4 composites. From Ref. [79]. Reprinted with permission from Elsevier.
Figure 9. Photocatalytic degradation efficiency of Methylene blue (MB) by graphitic carbon nitride (g-C3N4) and C/g-C3N4 composites. From Ref. [79]. Reprinted with permission from Elsevier.
Catalysts 09 00020 g009
Figure 10. TEM (a,c), HR-TEM (b), and SAED (d) images of double-shelled and sandwiched TiO2@Au@C hollow spheres. From Ref. [86]. Reprinted with permission from ACS.
Figure 10. TEM (a,c), HR-TEM (b), and SAED (d) images of double-shelled and sandwiched TiO2@Au@C hollow spheres. From Ref. [86]. Reprinted with permission from ACS.
Catalysts 09 00020 g010
Figure 11. Linear transform ln(C0/C) = f(t) of the kinetic curves of (a) 4-nitroaniline or (b) 4-nitrophenol photodegradation by P25, hollow TiO2, hollow TiO2@C, hollow TiO2/Au, and double-shelled and sandwiched TiO2@Au@C hollow spheres under visible light irradiation. From Ref. [86]. Reprinted with permission from ACS.
Figure 11. Linear transform ln(C0/C) = f(t) of the kinetic curves of (a) 4-nitroaniline or (b) 4-nitrophenol photodegradation by P25, hollow TiO2, hollow TiO2@C, hollow TiO2/Au, and double-shelled and sandwiched TiO2@Au@C hollow spheres under visible light irradiation. From Ref. [86]. Reprinted with permission from ACS.
Catalysts 09 00020 g011
Figure 12. (a) The reaction rate for methyl orange (MO) and phenol photodegradation under UV-VIS irradiation; (b) the recycling of MO photodegradation using ZEG-12; (c) photocurrent–voltage (I–V) curves of ZnO in dark and under irradiation; (d) transient photocurrent (I–t plots) at 0.85 V vs. RHE. From Ref. [91]. Reprinted with permission from Elsevier.
Figure 12. (a) The reaction rate for methyl orange (MO) and phenol photodegradation under UV-VIS irradiation; (b) the recycling of MO photodegradation using ZEG-12; (c) photocurrent–voltage (I–V) curves of ZnO in dark and under irradiation; (d) transient photocurrent (I–t plots) at 0.85 V vs. RHE. From Ref. [91]. Reprinted with permission from Elsevier.
Catalysts 09 00020 g012

Share and Cite

MDPI and ACS Style

Shi, Z.-J.; Ma, M.-G.; Zhu, J.-F. Recent Development of Photocatalysts Containing Carbon Species: A Review. Catalysts 2019, 9, 20. https://doi.org/10.3390/catal9010020

AMA Style

Shi Z-J, Ma M-G, Zhu J-F. Recent Development of Photocatalysts Containing Carbon Species: A Review. Catalysts. 2019; 9(1):20. https://doi.org/10.3390/catal9010020

Chicago/Turabian Style

Shi, Zheng-Jun, Ming-Guo Ma, and Jie-Fang Zhu. 2019. "Recent Development of Photocatalysts Containing Carbon Species: A Review" Catalysts 9, no. 1: 20. https://doi.org/10.3390/catal9010020

APA Style

Shi, Z. -J., Ma, M. -G., & Zhu, J. -F. (2019). Recent Development of Photocatalysts Containing Carbon Species: A Review. Catalysts, 9(1), 20. https://doi.org/10.3390/catal9010020

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