Next Article in Journal
The Effect of Different Oxygen Surface Functionalization of Carbon Nanotubes on the Electrical Resistivity and Strain Sensing Function of Cement Pastes
Next Article in Special Issue
Supported Ultra-Thin Alumina Membranes with Graphene as Efficient Interference Enhanced Raman Scattering Platforms for Sensing
Previous Article in Journal
A Nano-Rattle SnO2@carbon Composite Anode Material for High-Energy Li-ion Batteries by Melt Diffusion Impregnation
Previous Article in Special Issue
Recent Advancement of Electromagnetic Interference (EMI) Shielding of Two Dimensional (2D) MXene and Graphene Aerogel Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Noble Metal-Free TiO2-Coated Carbon Nitride Layers for Enhanced Visible Light-Driven Photocatalysis

National Engineering Research Center of Industry Crystallization Technology, School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2020, 10(4), 805; https://doi.org/10.3390/nano10040805
Submission received: 2 April 2020 / Revised: 21 April 2020 / Accepted: 21 April 2020 / Published: 23 April 2020

Abstract

:
Composites of g-C3N4/TiO2 were one-step prepared using electron impact with dielectric barrier discharge (DBD) plasma as the electron source. Due to the low operation temperature, TiO2 by the plasma method shows higher specific surface area and smaller particle size than that prepared via conventional calcination. Most interestingly, electron impact produces more oxygen vacancy on TiO2, which facilitates the recombination and formation of heterostructure of g-C3N4/TiO2. The composites have higher light absorption capacity and lower charge recombination efficiency. g-C3N4/TiO2 by plasma can produce hydrogen at a rate of 219.9 μmol·g−1·h−1 and completely degrade Rhodamine B (20mg·L−1) in two hours. Its hydrogen production rates were 3 and 1.5 times higher than that by calcination and pure g-C3N4, respectively. Electron impact, ozone and oxygen radical also play key roles in plasma preparation. Plasma has unique advantages in metal oxides defect engineering and the preparation of heterostructured composites with prospective applications as photocatalysts for pollutant degradation and water splitting.

Graphical Abstract

1. Introduction

Photocatalytic technology, a promising strategy for addressing energy shortages and environmental pollution, is important for the production of hydrogen via water splitting and the degradation of organic pollutants [1,2,3,4]. TiO2, discovered by Fujishima in 1972 [5], is the most widely studied and applied semiconductor photocatalyst and is non-toxic, stable, and cheap [6,7,8]. Many preparation methods of TiO2 and TiO2 composite have been used, such as the sol-gel method [9], solvothermal method [10], and chemical vapor deposition method [11] and so on. However, a green, simple, cheap and energy-efficient way for catalyst preparation is still necessary.
Non-thermal plasma, which has relatively low bulk temperature and extremely high electron temperature, has excellent advantages in preparing catalysts [12,13,14]. The catalyst can be prepared quickly without serious agglomeration due to the low temperature and high energy of plasma. On the other hand, the nucleation and crystallization process of the catalyst is very unique in plasma [13]. Furthermore, the catalysts prepared using the plasma method have small particles, strong interaction and specific structures [15,16]. However, TiO2 alone can only absorb ultraviolet light (only 4% of solar energy), even if the light absorption properties of TiO2 are improved. In addition, higher photo-generated charge recombination efficiency also affects its photocatalytic activity [17].
Strategies have been proposed to increase the photocatalytic activity of TiO2 under visible light. For example, compounding TiO2 with a narrow bandgap semiconductor catalyst can enhance its absorption of visible light and construct a special heterostructure [18,19]. Doping elements can reduce the TiO2 bandgap and increase its light absorption range [20,21]. Loading noble metals to TiO2 as the co-catalyst can act as its active site to enhance photocatalytic activity [22,23]. TiO2-coated carbon nitride layers as composites for enhanced photocatalytic activity is a more attractive approach. The g-C3N4 is widely used in the degradation of organic pollutants and water splitting by visible light irradiation due to its strong visible light response, high thermal resistance and chemical stability [24,25,26,27]. g-C3N4/TiO2 composites can not only transfer the photo-generated charge of g-C3N4 to TiO2 to increase its charge separation efficiency, but also reduce bandgap to increase its visible light absorption region [28,29,30]. Ma et al. got highly photocatalytic water splitting performance with g-C3N4/TiO2 composites by solvothermal method under visible light [10]. Papailias et al. utilized high temperature calcination to synthesize g-C3N4/TiO2 nanocomposites for NOx removal [31].
In this work, g-C3N4/TiO2 composites were prepared by using dielectric barrier discharge plasma, and their photocatalytic activities were evaluated by degrading RhB and hydrogen evolution under visible light irradiation. Due to the characteristics of the plasma preparation method, g-C3N4/TiO2 composites have many different properties compared to that by the traditional calcination method. Meanwhile, the mechanism of photocatalytic process and dielectric barrier discharge plasma preparation was proposed, respectively. It is predictable that the plasma method for catalysts preparation will be a very promising field.

2. Materials and Methods

2.1. Materials

Tetrabutyl titanate (TBT), melamine, absolute ethanol, rhodamine B (RhB), and triethanolamine were purchased from Shanghai Aladdin Biochemical Technology Co. (Shanghai, China). All chemicals were used directly.

Synthesis of TiO2 and g-C3N4/TiO2

g-C3N4 was dispersed by ultrasound in 20 mL of ethanol. We added 3ml TBT slowly into the as-prepared g-C3N4 suspension under adequate stirring. Then, the as-prepared hybrid suspensions were stand still for 24 h. The sample was filtered after 24 h. and drying at 80 °C. The samples were divided into two portions.
One portion was treated by DBD plasma to obtain g-C3N4/TiO2 composites according to the procedure reported in previous work [32,33]. Figure S1a shows the DBD device. The plasma was generated by the high voltage generator that can provide a sinusoidal waveform at 22 kHz with a voltage range of 0 to 30 kV. There are two quartz plates and a quartz ring between the two electrodes, and the sample was placed in the quartz ring. The gas atmosphere of the DBD reactor was air. The average power and average voltage of DBD during catalyst preparation were 200 W and 100 V, respectively. One-time plasma operation proceeded for 3 min to restrict the heat effect, followed by manually stirring to expose the untreated samples outside. The operation was repeated 20 times, until total plasma treatment time was 1 h. As shown in Figure S1b, the infrared (IR) image taken by the IR camera (Ircon, 100PHT, Everett, WA, USA) shown that the temperature of DBD plasma was below 106 °C. Finally, the obtained samples were denoted as TCNX-D (X = 10, 30, 50, 70, 90). X is equal to the weight ratio of g-C3N4 in the composites. D represents the samples prepared by DBD plasma. Figure S2 shows the schematic illustration of preparation of g-C3N4/TiO2 composite.
For comparison, another portion of the sample was calcined 450 °C at a rate of 5 °C/min for 2 h and designated as TCN50-C (50 wt% g-C3N4/TiO2).
The preparation method of pure TiO2 is the same as the above method. The samples obtained were denoted as TiO2-D and TiO2-C according to the different preparation methods.

2.2. Characterization

A Rigaku D/Max-2500 V/PC diffractometer with Cu Kα1 radiation (Cu Kα1 α = 0.154 nm, 40 kV, 40 mA, 8°·min−1, Rigaku, Tokyo, Japan) was used to analyze the crystal phase of samples. Fourier transform infrared (FTIR) spectra was analyzed by a Bruker Alpha FTIR-attenuated total reflection (ATR) instrument (Bruker, Karlsruhe, Germany). A Biaode SSA-7000 analyzer (Biaode Electronic Technology Ltd., Beijing, China) was applied to determine specific surface area and pore size. Thermogravimetry analysis was conducted on Perkin-Elmer TGA/DTA thermo-gravimetric analyzer (Waltham, MA, USA) in O2 atmosphere. The temperature range was 20–800 °C. An ULVAC-PHI-5000versaprobe instrument (Tokyo, Japan) was used to obtain X-ray photoelectron spectra. The adventitious carbon C1s used for element correction is located at 284.8 eV. Electron paramagnetic resonance measurement was obtained on a Bruker A300 spectrometer (Karlsruhe, Germany) at room temperature. The ultraviolet–visible (UV–vis) diffuse-reflectance spectroscopy was recorded by a UV-2600 spectrophotometer (Shimadzu, Kyoto, Japan). Photoluminescence (PL) spectroscopy was analyzed on Hitachi F-4600 spectrometer (Tokyo, Japan) with excitation at 350 nm. Scanning electron microscopy (SEM) and energy dispersive X-ray spectroscopy were conducted on Hitachi S4800 instrument (Tokyo, Japan). High-resolution transmission electron microscopy (HRTEM) was obtained using a JEM-2100F instrument (JEOL, Tokyo, Japan).

2.3. Measurement of Photocatalytic Activity

2.3.1. RhB degradation

We mixed 50 mg samples and 100 mL RhB (20 mg·L−1) solution uniformly in continuous stirring. First the mixture was stirred without light for 0.5 h to obtain the balance of adsorption and desorption. Visible light was provided by Xenon lamp (HSX-F300, 300 W) with a 420 nm cut filter. The illumination power was 100 mW/cm2. The light source was placed about 10 cm above the RhB solution. In the experiment, 4 mL solution was taken every 20 min for RhB degradation rate test. The degradation rate of the obtained RhB solution was determined using a UV-2600 instrument.

2.3.2. Hydrogen Generation

The H2 evolution experiment was conducted on the PerfectLight 3AG instrument (Beijing, China). The amount of H2 was determined by online gas chromatography. 100 mg sample was added to 70 mL of solution containing 10% triethanolamine and 133 μL of H2PtCl6 (0.5 wt%). The mixed solution was placed in 100 mL closed glass container. The 300 W xenon lamp was equipped with total reflection and 420 nm filter to provide full-spectrum light and visible light, respectively. The light source was placed 10 cm above the suspension. After full-spectrum irradiation for 3 h in vacuum, the Pt was loaded on the samples. Then the hydrogen production reaction started, and the content of H2 in the system was measured by online gas chromatography per hour.

2.3.3. Photoelectrochemical Investigation

The photocurrent density was conducted at the LK2010 electrochemical system. It is a traditional three-electrode system, tin oxide mixed with fluorine (FTO) conductive glass loaded with sample as working electrode. The working electrode was made as follows: 50 mg of the sample and 10 μL of nafion (5%) were added to 500 μL absolute ethanol, 200 μL of the suspension was loaded on FTO glass after mixing evenly. Then the FTO working electrode was obtained by drying at room temperature for 12 h. The reference electrode and counter electrode were the saturated Hg/HgO electrode and Pt wire, respectively. The 0.1M Na2SO4 solution was used as electrolyte. Visible light was provided by 300 W Xe lamp equipped with 420 nm filter. At the beginning of the test, we fully introduced N2 into the electrolyte to remove dissolved oxygen. The experimental potential is 0.4 eV, which is the optimal value by testing (see Figure S3).

3. Results and Discussion

3.1. Physico-Chemical Properties

The crystal phases of g-C3N4, TiO2, and g-C3N4/TiO2 composites were analyzed by using wide-angle X-ray diffraction (XRD). As shown in Figure 1a, g-C3N4 exhibits (100) and (002) planes at 13.2° and 27.6°, which correspond to tri-s-triazine structure and stacking of the conjugated aromatic system, respectively [34]. The as-prepared TiO2 present the diffraction peaks at 25.1°, 37.7°, 47.8°, 54.0°, 55.1° and 62.5°, which corresponded to the (101), (004), (200), (105), (211) and (204) crystal planes of anatase TiO2 (JCPDS 71-1166), respectively [35]. It indicated TiO2 was successfully prepared by DBD plasma. As the proportion of g-C3N4 increased in g-C3N4/TiO2 composites, the intensities of g-C3N4 peaks increased gradually. However, the no peak position shift of TiO2 indicates that g-C3N4 has no influence on the crystal structure of TiO2. Moreover, compared with samples (TiO2 and g-C3N4/TiO2 composites) prepared by high-temperature calcination, the samples by DBD plasma have lower peak intensity and wider peak width. On the basis of the full width at half maximum (FWHM) of (101) crystal plane and Debye–Scherrer equation, the average crystalline sizes of TiO2 in the g-C3N4/TiO2 composites were calculated and are listed in Table 1 [36]. The crystalline sizes of TiO2 in the composites decreased gradually as g-C3N4 increased. It is worth noting that TiO2 has smaller crystallite size than that by the calcination method. The relative low temperature of plasma is the main reason restricting the agglomeration of particles.
Fourier transform infrared spectra (FTIR) of g-C3N4, TiO2, and g-C3N4/TiO2 composites are shown in Figure 1b. It reveals the composition and chemical bonding of samples. For pure g-C3N4, the peak of 807 cm−1 is due to the tri-s-triazine unit structure, and four intense bands in the region 1240–1640 cm−1 are attributed to the stretching of the C–N heterocycle in g-C3N4 [37]. For TiO2-C and TiO2-D, they showed similar peaks, and the shoulder bands between 3000–3500 cm−1 can be attributed to –OH stretching vibration [38]. The shoulder bands near 3200 cm−1 in TCN50 composites are contributed by the N-H stretching vibration modes [37,39]. TCN50 composite exhibited both the g-C3N4 and TiO2 characteristic peaks. This shows that g-C3N4/TiO2 composites were directly prepared by DBD plasma and is agreement with XRD analysis results. During plasma process, high energy electron bombard samples. It can be deduced that the surface hydroxyl and amino groups enhance interaction between TiO2 and g-C3N4 [40]. The strong interfacial connection can be used as the channel for charge conduction to improve charge separation efficiency.
Nitrogen adsorption–desorption isotherms and pore-size distribution curves of g-C3N4, TiO2 and TCN50 samples are shown in Figure S4. It is clearly observed that samples prepared by plasma have larger N2 adsorption capacity. The specific data are listed in Table 1. The specific surface area of TiO2-D (64.56 m2/g) is approximately 2.3 times that of TiO2-C (28.18 m2/g), which can be ascribed to the low agglomeration of the sample’s particles prepared by low-temperature plasma. The specific surface area of g-C3N4 merely is 7.13 m2/g. As shown in Table S1, the specific surface area and pore volume of g-C3N4/TiO2 composites decreased and the average pore radius increased gradually as the proportion of g-C3N4 increased. Compared with TCN50-C, TCN50-D has higher specific surface area, pore volume, and smaller average pore radius, which indicates that the surface of the plasma-treated sample can expose more active sites. In addition to the low temperature, the repulsion between the electrons attached to the particles prevents the agglomerating of particles [12], resulting in the excellent dispersion of TiO2 on g-C3N4 to enhance photocatalytic activity.
Thermogravimetry (TG) was used to analyze thermal stability of the samples. As shown in Figure S5, the weight of TiO2 prepared by DBD plasma only reduced by 2%, which indicated that DBD plasma can complete decompose TBT to TiO2 at moderate temperature. The g-C3N4 gradually lost weight from 550 °C to 730 °C. The composites TCNX gradually lost weight from 520 °C to 650 °C, mainly due to the burning of g-C3N4. The thermal stability of g-C3N4 was reduced after coating with TiO2. The reason is catalytic action of TiO2 and the cross-linking ring of g-C3N4 after compounding [41]. Ignoring the slight weight loss due to water, the actual proportions of g-C3N4 in the composites TCNX-D (X = 10, 30, 50, 50, 70, 90) were 13.0, 30.3, 52.9, 71.3, and 90.0 wt%, respectively. The proportion of g-C3N4 in TCN50-C was 53.6%, which was approximately equal to the ideal ratio of TCN50-D.

3.2. Characterizations of Oxygen Vacancies

Chemical states of elements in g-C3N4, TiO2, and g-C3N4/TiO2 composites were investigated by X-ray photoelectron spectroscopy (XPS). Figure S6a shows the high-resolution C 1s spectrum of the prepared samples. All samples have two C 1s peaks at 284.8 and 288.3 eV, which can be ascribed to the inherent adventitious carbon and N-C-N coordination, respectively [42]. Compared with the N-C-N peak of pure g-C3N4, TCN50 exhibit weaker peak intensity and positive shifts of 0.4 eV in binding energies, which indicates there is a chemical interaction between TiO2 and g-C3N4, hence, g-C3N4 has close surface contact with TiO2 [43]. As shown in Figure S5b, for the N 1 s high-resolution spectrum of pure g-C3N4 and g-C3N4/TiO2 composites, three peaks were observed at about 398.8, 399.6, and 401.2 eV. The first peak is attributed to sp2 hybridized nitrogen (C=N–C), the second peak is due to the tertiary N in N–(C)3 groups, and the last peak corresponds to the existence of amino groups (C–N–H) [44,45].
Figure 2a,b shows O 1s and Ti 2p spectrum region of pure TiO2. Figure 2a shows the O 1s spectrum region of TiO2-D have three peaks at about 529.7, 532.0 and 533.4 eV, which corresponds to the lattice oxygen, the oxygen vacancy and the adsorbed oxygen, respectively [46,47]. By calculating the ratio of the area occupied by oxygen vacancies, it was 23.4% and 18.6% in TiO2-D and TiO2-C, respectively, indicating that plasma method can produce more oxygen vacancies than that of the calcination method in TiO2. Figure 2b shows the Ti 2p spectrum of pure TiO2, There are four peaks at about 458.2, 463.7, 458.9, and 464.6 eV, corresponding to Ti3+ 2p3/2, Ti3+ 2p1/2, Ti4+ 2p3/2 and Ti4+ 2p1/2, respectively [46,47]. It was found that the area of Ti3+ in TiO2-D and TiO2-C were 15.6% and 11.1%, respectively. Compared with TiO2 by calcination, the valence of more Ti in TiO2 by plasma is reduced from Ti4+ to Ti3+. This is consistent with the observation for the oxygen vacancy. During the plasma preparation of TiO2, the oxygen atoms escaped to form oxygen vacancies and trivalent Ti [46].
Figure 2c,d shows O 1s and Ti 2p spectrum of TCN50 composites. TCN50 and TiO2 show similar peaks at O 1s and Ti 2p. However, the intensity of the peaks representing oxygen vacancies and Ti3+ decreased. The ratio of oxygen vacancies and Ti3+ were 23.4% and 11.2% in TCN50-D, 18.6% and 7.2% in TCN50-C, respectively. It is indicated that g-C3N4 occupies the oxygen vacancies of TiO2 after TiO2 coated on g-C3N4. This will undoubtedly strengthen the interaction of TiO2 and g-C3N4. It is in agreement with previous results.
To further verify the existence of oxygen vacancies, electron paramagnetic resonance (EPR) analysis of the TCN50 composites was performed. As shown in Figure 3a, both TCN50-C and TCN50-D have the EPR signal at g = 2.003, which means the appearance of oxygen vacancies [48]. The peak intensity of TCN50-D is stronger than TCN50-C, indicating the plasma method produces more oxygen vacancies. As can be seen from Figure S7, the EPR spectra of TiO2-C and D exhibit the same characteristics as that of TCN50. This is consistent with XPS results. In addition, Figure 3b shows different colors between TiO2-D and TiO2-C. The color of TiO2-C is white, while TiO2-D is gray. This indicates oxygen vacancies can narrow band gap to promote light harvesting, leading significant color change [49].

3.3. Optical Properties

Ultraviolet–visible diffuse reflection spectrum (UV-vis DRS) is shown in Figure 4a and Figure S8 reveal the light absorption capacity and bandgaps of the as-prepared g-C3N4, TiO2, and TCNX. The critical values of the light response of TiO2-D and TiO2-C were found at 398 nm and 388 nm, indicating that bandgaps were 3.12 and 3.19 eV, respectively [50]. The absorption starting point of g-C3N4 was located at 458 nm. The absorption range of visible light of TCN50 composites was significantly expanded due to introducing g-C3N4. As shown in Figure 4b, on the basis of the Kubelka-Munk formula, the bandgaps of g-C3N4, TCN50-C, and TCN50-D were calculated to be 2.71, 2.70, and 2.63 eV, respectively. Based on the above results, the samples prepared by plasma had a narrower bandgap. This can be attributed to oxygen vacancies [51]. Oxygen vacancies can introduce a defect status below the conduction band, and thus narrows the bandgap to improve the light absorption range [46,52]. Hence, TCN50-D has a relatively strong visible light response capacity.
Photoluminescence (PL) spectra was shown in Figure 4c. The PL peak intensity is proportional to the recombination rates of the photogenerated electron-hole pairs [53]. g-C3N4 has strong peak intensity, indicating that its charge separation efficiency is low. TiO2-D has weaker PL peak strength than that of TiO2-C, indicating that TiO2-D is more conducive to charge transport. After TiO2 coated on g-C3N4, PL peak intensity was greatly reduced, because electrons can be transmitted through the interface of g-C3N4 and TiO2 [54]. TCN50-D shows lower peak intensity than TCN50-C, resulting in that TCN50-D has lower charge recombination rate and stronger photocatalytic activity. Meanwhile, the presence of oxygen vacancies can promote the separation of electrons and holes [55]. The oxygen vacancies not only restrain the recombination of charges, but also narrow the band gap to increase light absorption [52,56]. As shown in Figure 4d, TCN50-D exhibits the highest photocurrent density than TCN50-C, indicating that TCN50-D has higher photoelectric conversion efficiency [57].

3.4. Morphologies

Figure 5 shows the morphology and microstructure of TiO2, g-C3N4 and g-C3N4/TiO2 composites. As shown in Figure 5a,b, TiO2-D has less particle aggregation than TiO2-C, which agreed well with previous result. As shown in Figure 5c,d, g-C3N4 has an anomalous layered structure and smooth flat surface. After TiO2 coating process by plasma, TiO2 was uniformly dispersed on the surface of g-C3N4, and the surface of g-C3N4 changes from smooth to rough. Figure 5e and Figure S9 show the energy dispersive X-ray spectroscopy (EDS) mapping of TCN50-D, the four elements (C, N, O, Ti) are well-dispersed in the TCN50-D composite. It can form a heterostructure for electron transport and reduce charge recombination efficiency.
To further investigate the microstructure of g-C3N4/TiO2 composites, HRTEM was conducted. As can be seen from Figure 5f,g, the particle size of TiO2 in TCN50-C composite is around 18 nm, that in the TCN50-D composite is around 12 nm. This proves that plasma treated sample has smaller particle size. The lattice spacing of TiO2 is 0.35 nm representing the (101) plane of anatase titanium oxide. Figure 5h also shows the clear lattice fringe of g-C3N4 is 0.32 nm, corresponding to the (002) lattice plane of g-C3N4 [54]. Compared with TCN50-C composite, TCN50-D have more and tighter heterojunctions, which also means extensive interfacial contacts to enhance photocatalytic activity.

3.5. Photocatalytic Activity

RhB degradation and H2 production under visible light were conducted to evaluate the photocatalytic performance of the samples. As shown in Figure 6a, the concentration of RhB hardly changed within 0.5 h of dark adsorption. For TiO2-C and TiO2-D, RhB was degraded by 14.3% and 33.7% under two hours of the irradiation of visible light, respectively, which can be ascribed to the sensitization of dye [58]. Plasma-treated TiO2 show better photoactivity. Pure g-C3N4 degraded 63.7% of RhB under two hours of visible light irradiation. With the increase of proportion of g-C3N4 in g-C3N4/TiO2, the degradation efficiency increased first and then decreased; 50 wt% g-C3N4 in g-C3N4/TiO2 composite exhibited the highest RhB degradation efficiency. When the amount of g-C3N4 is low, TiO2 covers its surface and prevents absorption of visible light. However, when the amount of TiO2 is low, it is insufficient to promote charge separation. Therefore, the appropriate ratio of g-C3N4/TiO2 has higher photocatalytic activity. Compared with 79.7% RhB degradation efficiency of TCN50-C, TCN50-D can completely degrade RhB in two hours, due to its stronger visible light absorption and charge separation efficiency.
Furthermore, Figure 6b shows the reusability experiment of the TCN50-D sample. The photocatalytic activity for degradation of RhB decreased from 95.5% to 90.2% within 6 h under visible light after the third cycle. This indicated that the g-C3N4/TiO2 composite by plasma has high stability.
As shown in Figure 6c, the hydrogen production of pure TiO2 was not detected in the system. Pure g-C3N4 only produced 72.1 μmol·g−1 H2 per hour. The hydrogen production of TCN50-D and TCN50-C composites was 219.9 and 174.3 μmol·g−1·h−1, respectively. TCN50-D exhibited higher photocatalytic activity and approximately 3 times that of pure g-C3N4 and 1.26 times that of TCN50-C.

3.6. Mechanism

In order to investigate the phtotcatalytic degradation mechanism of the TCN50-D composite, different scavengers were added before RhB degraded by TCN50-D. As shown in the Figure S10, the degradation activity of TCN50-D changed slightly after the addition of the hydroxyl radical scavenger isopropyl alcohol (IPA) and the superoxide radical scavenger p-benzoquinone (BQ). However, after the addition of the hole scavenger Ethylene Diamine Tetraacetic Acid-2Na (EDTA-2Na), a significant decrease in photocatalytic activity for RhB degradation was observed, indicating that the hole is the main active species for degrading RhB. By measuring the intersection of the slope of the XPS valence band curve and the X axis, the valence band (VB) of the samples can be determined. As shown in Figure S11, the VB of TCN50-C and TCN50-D is 2.32 eV and 2.38 eV, respectively, indicating that TCN50-D possesses stronger oxidizing ability [40]. The VB of TiO2-D and TiO2-C is 2.91 eV and 2.84 eV, respectively. The conduction band (CB) of TiO2-D can be calculated that is −0.20 eV, according to its band gap of 3.11 eV. The VB and CB of g-C3N4 are 2.24 eV and −0.47 eV, respectively [40]. Therefore, the energy band structure model of TCN50-D can be constructed and shown in Figure 6d. Under visible light irradiation, g-C3N4 captures photons to generate electron-hole pairs, then electrons can be transported from the CB of g-C3N4 to the CB of TiO2-D to reduce H+ to produce H2. Meanwhile, the holes in the VB of g-C3N4 can oxidize RhB to exert photodegradation [59].
The mechanism of TiO2 and g-C3N4/TiO2 composite prepared by dielectric barrier discharge plasma is proposed as follows. The high-energy electron bombardment plays a key role in the process.
During the plasma process, there are a large number of electrons that can decompose titanium hydroxide through breaking bonds by electron bombardment (Equation (1)). This also can produce oxygen vacancies.
Ti(OH)4 + e → TiO2 + H2O
Under the influence of plasma, the ozone and oxygen atoms can be generated by using Equations (2)–(4) [16,60].
e + O2 → e + O(3P) + O(3P)
e + O2 → e + O(1D) + O(3P)
O + O2 + M → O3* +M → O3 + M
As shown in Equations (5) and (6), the oxygen atom and ozone can also lead to Ti(OH)4 decomposition.
Ti(OH)4 + O → TiO2 + 2H2O + 1/2O2
Ti(OH)4 + O3 → TiO2 + 2H2O + 3/2O2

4. Conclusions

TiO2 and g-C3N4/TiO2 composites were prepared by DBD plasma, which is a green, easy, and efficient method for catalyst preparation. Here, TiO2 by plasma has enriched oxygen vacancies and larger specific surface area due to the electron impact. The electron impact, ozone and oxygen radical play important role in plasma preparation, which facilitate the interaction of TiO2 and g-C3N4 and forms heterojunctions. g-C3N4/TiO2 composites by plasma have stronger light absorption capacity and higher charge separation efficiency. TCN50-D exhibited the highest photocatalysis activity on RhB degradation and hydrogen production, which was ascribed to enriched oxygen vacancies and special heterostructures. Plasma, green and convenient technology, provides a promising strategy for oxide defect engineering and composite preparation.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/4/805/s1: Figure S1: The dielectric barrier discharge (DBD) plasma reactor and infrared (IR) image of DBD plasma, Figure S2: The schematic illustration for the preparation of TCN composites photocatalysts, Figure S3: Photocurrent tests of TCN50-D at different potential, Figure S4: Nitrogen adsorption–desorption isotherms (a) and pore-size distribution curves (b) of g-C3N4, TiO2 and TCN50 samples, Figure S5: Thermogravimetric analysis curves, Figure S6: The high-resolution XPS spectra C 1s and N 1s, Figure S7: The EPR spectra of TiO2-C and D samples, Figure S8: FTIR spectra, UV–vis DRS and XPS valence band (VB) spectra of TCNX, Figure S9: EDS patterns and element mapping of TCN50-D sample, Figure S10: Trapping experimental of photogenerated radicals and holes in TCN50-D sample for the RhB degradation, Figure S11: VB XPS of TiO2-C, TiO2-D, TCN50-C and TCN50-D, Table S1: Surface and structural characterization of TiO2, g-C3N4 and TCNX composites.

Author Contributions

Z.W. conceived the experiments. B.Z. conducted the preparation and photocatalysis experiment. X.P. set up the experimental device. B.Z. and Z.W. wrote the original draft of the manuscript. All authors contributed to the discussion of the results. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Key Research and Development Program of China (No.2016YFF0102503), National Natural Science Foundation of China (No.21878214) and State Key Laboratory of Efficient Utilization for Low Grade Phosphate Rock and Its Associated Resources (No.WFKF2019-03).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chava, R.K.; Son, N.; Kim, Y.S.; Kang, M. Controlled Growth and Bandstructure Properties of One Dimensional Cadmium Sulfide Nanorods for Visible Photocatalytic Hydrogen Evolution Reaction. Nanomaterials 2020, 10, 619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Pan, Y.; You, Y.; Xin, S.; Li, Y.; Fu, G.; Cui, Z.; Men, Y.; Cao, F.; Yu, S.; Goodenough, J.B. Photocatalytic CO2 reduction by carbon-coated indium-oxide nanobelts. J. Am. Chem. Soc. 2017, 139, 4123–4129. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, Z.; Li, C.; Domen, K. Recent developments in heterogeneous photocatalysts for solar-driven overall water splitting. Chem. Soc. Rev. 2019, 48, 2109–2125. [Google Scholar] [CrossRef] [PubMed]
  4. Lin, L.; Yu, Z.; Wang, X. Crystalline carbon nitride semiconductors for photocatalytic water splitting. Angew. Chem. Int. Ed. 2019, 131, 6225–6236. [Google Scholar] [CrossRef]
  5. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37. [Google Scholar] [CrossRef]
  6. Nawaz, R.; Kait, C.F.; Chia, H.Y.; Isa, M.H.; Huei, L.W. Glycerol-mediated facile synthesis of colored titania nanoparticles for visible light photodegradation of phenolic compounds. Nanomaterials 2019, 9, 1586. [Google Scholar] [CrossRef] [Green Version]
  7. Yu, X.; Fan, X.; An, L.; Liu, G.; Li, Z.; Liu, J.; Hu, P. Mesocrystalline Ti3+-TiO2 hybridized g-C3N4 for efficient visible-light photocatalysis. Carbon 2018, 128, 21–30. [Google Scholar] [CrossRef]
  8. Xing, Z.; Zhang, J.; Cui, J.; Yin, J.; Zhao, T.; Kuang, J.; Xiu, Z.; Wan, N.; Zhou, W. Recent advances in floating TiO2-based photocatalysts for environmental application. Appl. Catal. B-Environ. 2018, 225, 452–467. [Google Scholar] [CrossRef]
  9. Moongraksathum, B.; Chen, Y. Anatase TiO2 co-doped with silver and ceria for antibacterial application. Catal. Today 2018, 310, 68–74. [Google Scholar] [CrossRef]
  10. Ma, J.; Tan, X.; Jiang, F.; Yu, T. Graphitic C3N4 nanosheet-sensitized brookite TiO2 to achieve photocatalytic hydrogen evolution under visible light. Catal. Sci. Technol. 2017, 7, 3275–3282. [Google Scholar] [CrossRef]
  11. Tan, Y.; Shu, Z.; Zhou, J.; Li, T.; Wang, W.; Zhao, Z. One-step synthesis of nanostructured g-C3N4/TiO2 composite for highly enhanced visible-light photocatalytic H2 evolution. Appl. Catal. B-Environ. 2018, 230, 260–268. [Google Scholar] [CrossRef]
  12. Li, J.; Ma, C.; Zhu, S.; Yu, F.; Dai, B.; Yang, D. A review of recent advances of dielectric barrier discharge plasma in catalysis. Nanomaterials 2019, 9, 1428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wang, Z.; Zhang, Y.; Neyts, E.C.; Cao, X.; Zhang, X.; Jang, B.W.L.; Liu, C. Catalyst preparation with plasmas: How does it work? ACS Catal. 2018, 8, 2093–2110. [Google Scholar] [CrossRef]
  14. Dou, S.; Tao, L.; Wang, R.; El Hankari, S.; Chen, R.; Wang, S. Plasma-assisted synthesis and surface modification of electrode materials for renewable energy. Adv. Mater. 2018, 30, 1705850. [Google Scholar] [CrossRef]
  15. Yan, X.; Zhao, B.; Liu, Y.; Li, Y. Dielectric barrier discharge plasma for preparation of Ni-based catalysts with enhanced coke resistance: Current status and perspective. Catal. Today 2015, 256, 29–40. [Google Scholar] [CrossRef]
  16. Sun, Q.; Yu, B.; Liu, C. Characterization of ZnO nanotube fabricated by the plasma decomposition of Zn(OH)2 via dielectric barrier discharge. Plasma Chem. Plasma Process. 2012, 32, 201–209. [Google Scholar] [CrossRef]
  17. Zada, A.; Qu, Y.; Ali, S.; Sun, N.; Lu, H.; Yan, R.; Zhang, X.; Jing, L. Improved visible-light activities for degrading pollutants on TiO2/g-C3N4 nanocomposites by decorating SPR Au nanoparticles and 2,4-dichlorophenol decomposition path. J. Hazard. Mater. 2018, 342, 715–723. [Google Scholar] [CrossRef]
  18. Tahir, M.B.; Sagir, M.; Shahzad, K. Removal of acetylsalicylate and methyl-theobromine from aqueous environment using nano-photocatalyst WO3-TiO2 @g-C3N4 composite. J. Hazard. Mater. 2019, 363, 205–213. [Google Scholar] [CrossRef]
  19. Zhao, W.; Yang, X.; Liu, C.; Qian, X.; Wen, Y.; Yang, Q.; Sun, T.; Chang, W.; Liu, X.; Chen, Z. Facile construction of all-solid-state Z-scheme g-C3N4/TiO2 thin film for the efficient visible-light degradation of organic pollutant. Nanomaterials 2020, 10, 600. [Google Scholar] [CrossRef] [Green Version]
  20. Marques, J.; Gomes, T.D.; Forte, M.A.; Silva, R.F.; Tavares, C.J. A new route for the synthesis of highly-active N-doped TiO2 nanoparticles for visible light photocatalysis using urea as nitrogen precursor. Catal. Today 2019, 326, 36–45. [Google Scholar] [CrossRef]
  21. Lu, Z.; Zeng, L.; Song, W.; Qin, Z.; Zeng, D.; Xie, C. In situ synthesis of C-TiO2/g-C3N4 heterojunction nanocomposite as highly visible light active photocatalyst originated from effective interfacial charge transfer. Appl. Catal. B-Environ. 2017, 202, 489–499. [Google Scholar] [CrossRef]
  22. Sanabria-Arenas, B.E.; Mazare, A.; Yoo, J.; Nhat, T.N.; Hejazi, S.; Bian, H.; Diamanti, M.V.; Pedeferri, M.P.; Schmuki, P. Intrinsic AuPt-alloy particles decorated on TiO2 nanotubes provide enhanced photocatalytic degradation. Electrochim. Acta 2018, 292, 865–870. [Google Scholar] [CrossRef]
  23. Hampel, B.; Kovacs, G.; Czekes, Z.; Hernadi, K.; Danciu, V.; Ersen, O.; Girleanu, M.; Focsan, M.; Baia, L.; Pap, Z. Mapping the photocatalytic activity and ecotoxicology of Au, Pt/TiO2 composite photocatalysts. ACS Sustain. Chem. Eng. 2018, 6, 12993–13006. [Google Scholar] [CrossRef]
  24. 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]
  25. Yu, S.; Wang, Y.; Sun, F.; Wang, R.; Zhou, Y. Novel mpg-C3N4/TiO2 nanocomposite photocatalytic membrane reactor for sulfamethoxazole photodegradation. Chem. Eng. J. 2018, 337, 183–192. [Google Scholar] [CrossRef]
  26. Lee, G.; Lee, X.; Lyu, C.; Liu, N.; Andandan, S.; Wu, J.J. Sonochemical synthesis of copper-doped BiVO4/g-C3N4 nanocomposite materials for photocatalytic degradation of bisphenol a under simulated sunlight irradiation. Nanomaterials 2020, 10, 498. [Google Scholar] [CrossRef] [Green Version]
  27. Wang, Z.; Peng, X.; Tian, S.; Wang, Z. Enhanced hydrogen production from water on Pt/g-C3N4 by room temperature electron reduction. Mater. Res. Bull. 2018, 104, 1–5. [Google Scholar] [CrossRef]
  28. Yan, J.; Wu, H.; Chen, H.; Zhang, Y.; Zhang, F.; Liu, S.F. Fabrication of TiO2/C3N4 heterostructure for enhanced photocatalytic Z-scheme overall water splitting. Appl. Catal. B-Environ. 2016, 191, 130–137. [Google Scholar] [CrossRef]
  29. Liu, H.; Zhang, Z.; He, H.; Wang, X.; Zhang, J.; Zhang, Q.; Tong, Y.; Liu, H.; Ramakrishna, S.; Yan, S.; et al. One-step synthesis heterostructured g-C3N4/TiO2 composite for rapid degradation of pollutants in utilizing visible light. Nanomaterials 2018, 8, 842. [Google Scholar] [CrossRef] [Green Version]
  30. Smýkalová, A.; Sokolová, B.; Foniok, K.; Matějka, V.; Praus, P. Photocatalytic degradation of selected pharmaceuticals using g-C3N4 and TiO2 nanomaterials. Nanomaterials 2019, 9, 1194. [Google Scholar] [CrossRef] [Green Version]
  31. Papailias, I.; Todorova, N.; Giannakopoulou, T.; Yu, J.; Dimotikali, D.; Trapalis, C. Photocatalytic activity of modified g-C3N4 /TiO2 nanocomposites for NOx removal. Catal. Today 2017, 280, 37–44. [Google Scholar] [CrossRef]
  32. Zhang, B.; Wang, Z.; Peng, X.; Wang, Z.; Zhou, L.; Yin, Q. A novel route to manufacture 2D layer MoS2 and g-C3N4 by atmospheric plasma with enhanced visible-light-driven photocatalysis. Nanomaterials 2019, 9, 1139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Peng, X.; Wang, Z.; Wang, Z.; Gong, J.; Hao, H. Electron reduction for the preparation of rGO with high electrochemical activity. Catal. Today 2019, 337, 63–68. [Google Scholar] [CrossRef]
  34. Yang, H.; Jin, Z.; Hu, H.; Bi, Y.; Lu, G. Ni-Mo-S nanoparticles modified graphitic C3N4 for efficient hydrogen evolution. Appl. Surf. Sci. 2018, 427, 587–597. [Google Scholar] [CrossRef]
  35. Du, X.; Bai, X.; Xu, L.; Yang, L.; Jin, P. Visible-light activation of persulfate by TiO2/g-C3N4 photocatalyst toward efficient degradation of micropollutants. Chem. Eng. J. 2020, 384, 123245. [Google Scholar] [CrossRef]
  36. Hao, R.; Wang, G.; Jiang, C.; Tang, H.; Xu, Q. In situ hydrothermal synthesis of g-C3N4/TiO2 heterojunction photocatalysts with high specific surface area for Rhodamine B degradation. Appl. Surf. Sci. 2017, 411, 400–410. [Google Scholar] [CrossRef]
  37. Di, T.; Zhu, B.; Cheng, B.; Yu, J.; Xu, J. A direct Z-scheme g-C3N4/SnS2 photocatalyst with superior visible-light CO2 reduction performance. J. Catal. 2017, 352, 532–541. [Google Scholar] [CrossRef]
  38. Lei, J.; Chen, Y.; Shen, F.; Wang, L.; Liu, Y.; Zhang, J. Surface modification of TiO2 with g-C3N4 for enhanced UV and visible photocatalytic activity. J. Alloys Compd. 2015, 631, 328–334. [Google Scholar] [CrossRef]
  39. Wang, H.; Liang, Y.; Liu, L.; Hu, J.; Cui, W. Highly ordered TiO2 nanotube arrays wrapped with g-C3N4 nanoparticles for efficient charge separation and increased photoelectrocatalytic degradation of phenol. J. Hazard. Mater. 2018, 344, 369–380. [Google Scholar] [CrossRef]
  40. Ma, J.; Tan, X.; Yu, T.; Li, X. Fabrication of g-C3N4/TiO2 hierarchical spheres with reactive {001} TiO2 crystal facets and its visible-light photocatalytic activity. Int. J. Hydrogen Energy 2016, 41, 3877–3887. [Google Scholar] [CrossRef]
  41. Tong, Z.; Yang, D.; Xiao, T.; Tian, Y.; Jiang, Z. Biomimetic fabrication of g-C3N4/TiO2 nanosheets with enhanced photocatalytic activity toward organic pollutant degradation. Chem. Eng. J. 2015, 260, 117–125. [Google Scholar] [CrossRef]
  42. Ma, L.; Wang, G.; Jiang, C.; Bao, H.; Xu, Q. Synthesis of core-shell TiO2@g-C3N4 hollow microspheres for efficient photocatalytic degradation of rhodamine B under visible light. Appl. Surf. Sci. 2018, 430, 263–272. [Google Scholar] [CrossRef]
  43. Li, F.T.; Liu, S.J.; Xue, Y.B.; Wang, X.J.; Hao, Y.J.; Zhao, J.; Liu, R.H.; Zhao, D.S. Structure modification function of g-C3N4 for Al2O3 in the in situ hydrothermal process for enhanced photocatalytic activity. Chem. A Eur. J. 2015, 21, 10149–10159. [Google Scholar] [CrossRef]
  44. Wu, F.; Li, X.; Liu, W.; Zhang, S. Highly enhanced photocatalytic degradation of methylene blue over the indirect all-solid-state Z-scheme g-C3N4-RGO-TiO2 nanoheterojunctions. Appl. Surf. Sci. 2017, 405, 60–70. [Google Scholar] [CrossRef]
  45. Shi, L.; Yang, L.; Zhou, W.; Liu, Y.; Yin, L.; Hai, X.; Song, H.; Ye, J. Photoassisted construction of holey defective g-C3N4 photocatalysts for efficient visible-light-driven H2O2 production. Small 2018, 14, 1703142. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, F.; Feng, G.; Hu, M.; Huang, Y.; Zeng, H. Liquid-plasma hydrogenated synthesis of gray titania with engineered surface defects and superior photocatalytic activity. Nanomaterials 2020, 10, 342. [Google Scholar] [CrossRef] [Green Version]
  47. Shi, H.; Long, S.; Hu, S.; Hou, J.; Ni, W.; Song, C.; Li, K.; Gurzadyan, G.G.; Guo, X. Interfacial charge transfer in 0D/2D defect-rich heterostructures for efficient solar-driven CO2 reduction. Appl. Catal. B-Environ. 2019, 245, 760–769. [Google Scholar] [CrossRef]
  48. Feng, X.; Wang, P.; Hou, J.; Qian, J.; Ao, Y.; Wang, C. Significantly enhanced visible light photocatalytic efficiency of phosphorus doped TiO2 with surface oxygen vacancies for ciprofloxacin degradation: Synergistic effect and intermediates analysis. J. Hazard. Mater. 2018, 351, 196–205. [Google Scholar] [CrossRef]
  49. Zhang, Y.C.; Afzal, N.; Pan, L.; Zhang, X.; Zou, J.J. Structure-activity relationship of defective metal-based photocatalysts for water splitting: Experimental and theoretical perspectives. Adv. Sci. 2019, 6, 1900053. [Google Scholar] [CrossRef] [Green Version]
  50. Chen, X.; Wei, J.; Hou, R.; Liang, Y.; Xie, Z.; Zhu, Y.; Zhang, X.; Wang, H. Growth of g-C3N4 on mesoporous TiO2 spheres with high photocatalytic activity under visible light irradiation. Appl. Catal. B-Environ. 2016, 188, 342–350. [Google Scholar] [CrossRef]
  51. Li, Y.; Wang, W.; Wang, F.; Di, L.; Yang, S.; Zhu, S.; Yao, Y.; Ma, C.; Dai, B.; Yu, F. Enhanced photocatalytic degradation of organic dyes via defect-rich TiO2 prepared by dielectric barrier discharge plasma. Nanomaterials 2019, 9, 720. [Google Scholar] [CrossRef] [Green Version]
  52. Lv, Y.; Pan, C.; Ma, X.; Zong, R.; Bai, X.; Zhu, Y. Production of visible activity and UV performance enhancement of ZnO photocatalyst via vacuum deoxidation. Appl. Catal. B-Environ. 2013, 138–139, 26–32. [Google Scholar] [CrossRef]
  53. Xue, K.; He, R.; Yang, T.; Wang, J.; Sun, R.; Wang, L.; Yu, X.; Omeoga, U.; Pi, S.; Yang, T.; et al. MOF-based In2S3-X2S3 (X=Bi; Sb)@TFPT-COFs hybrid materials for enhanced photocatalytic performance under visible light. Appl. Surf. Sci. 2019, 493, 41–54. [Google Scholar] [CrossRef]
  54. Jiang, X.; Xing, Q.; Luo, X.; Li, F.; Zou, J.; Liu, S.; Li, X.; Wang, X. Simultaneous photoreduction of Uranium(VI) and photooxidation of Arsenic(III) in aqueous solution over g-C3N4/TiO2 heterostructured catalysts under simulated sunlight irradiation. Appl. Catal. B-Environ. 2018, 228, 29–38. [Google Scholar] [CrossRef]
  55. Wang, Y.; Cai, J.; Wu, M.; Chen, J.; Zhao, W.; Tian, Y.; Ding, T.; Zhang, J.; Jiang, Z.; Li, X. Rational construction of oxygen vacancies onto tungsten trioxide to improve visible light photocatalytic water oxidation reaction. Appl. Catal. B-Environ. 2018, 239, 398–407. [Google Scholar] [CrossRef]
  56. Li, J.; Weng, B.; Cai, S.; Chen, J.; Jia, H.; Xu, Y. Efficient promotion of charge transfer and separation in hydrogenated TiO2/WO3 with rich surface-oxygen-vacancies for photodecomposition of gaseous toluene. J. Hazard. Mater. 2018, 342, 661–669. [Google Scholar] [CrossRef]
  57. Deng, Y.; Tang, L.; Zeng, G.; Wang, J.; Zhou, Y.; Wang, J.; Tang, J.; Wang, L.; Feng, C. Facile fabrication of mediator-free Z-scheme photocatalyst of phosphorous-doped ultrathin graphitic carbon nitride nanosheets and bismuth vanadate composites with enhanced tetracycline degradation under visible light. J. Colloid Interface Sci. 2018, 509, 219–234. [Google Scholar] [CrossRef]
  58. Chang, F.; Zhang, J.; Xie, Y.; Chen, J.; Li, C.; Wang, J.; Luo, J.; Deng, B.; Hu, X. Fabrication, characterization, and photocatalytic performance of exfoliated g-C3N4-TiO2 hybrids. Appl. Surf. Sci. 2014, 311, 574–581. [Google Scholar] [CrossRef]
  59. Yu, J.; Wang, S.; Low, J.; Xiao, W. Enhanced photocatalytic performance of direct Z-scheme g-C3N4-TiO2 photocatalysts for the decomposition of formaldehyde in air. Phys. Chem. Chem. Phys. 2013, 15, 16883. [Google Scholar] [CrossRef]
  60. Kogelschatz, U. Dielectric-barrier discharges: Their history, discharge physics, and industrial applications. Plasma Chem. Plasma Process. 2003, 23, 1–46. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of TiO2, g-C3N4 and TCNX samples, (b) FTIR spectra of TiO2, g-C3N4, TCN50 samples.
Figure 1. (a) XRD patterns of TiO2, g-C3N4 and TCNX samples, (b) FTIR spectra of TiO2, g-C3N4, TCN50 samples.
Nanomaterials 10 00805 g001
Figure 2. The high-resolution XPS spectra: (a) O 1s, (b) Ti 2p, of TiO2 samples. (c) O 1s, (d) Ti 2p of TCN50 samples.
Figure 2. The high-resolution XPS spectra: (a) O 1s, (b) Ti 2p, of TiO2 samples. (c) O 1s, (d) Ti 2p of TCN50 samples.
Nanomaterials 10 00805 g002
Figure 3. (a) EPR spectra of TCN50-C and TCN50-D samples, (b) images of TiO2 prepared by plasma and calcination.
Figure 3. (a) EPR spectra of TCN50-C and TCN50-D samples, (b) images of TiO2 prepared by plasma and calcination.
Nanomaterials 10 00805 g003
Figure 4. (a) UV–vis DRS of TiO2, g-C3N4, TCN50-C and TCN50-D samples. (b) The relationship between (ahv)2 and photo energy. (c) Photoluminescence spectra of g-C3N4, TiO2 and TCN50 samples. (d) Photocurrent density vs. time for g-C3N4/FTO, TCN50-C/FTO and TCN50-D/FTO.
Figure 4. (a) UV–vis DRS of TiO2, g-C3N4, TCN50-C and TCN50-D samples. (b) The relationship between (ahv)2 and photo energy. (c) Photoluminescence spectra of g-C3N4, TiO2 and TCN50 samples. (d) Photocurrent density vs. time for g-C3N4/FTO, TCN50-C/FTO and TCN50-D/FTO.
Nanomaterials 10 00805 g004aNanomaterials 10 00805 g004b
Figure 5. SEM images of (a) TiO2-C, (b) TiO2-D, (c) g-C3N4, (d) TCN50-D and (e) EDS elemental mappings of TCN50-D samples; high-resolution transmission electron microscopy (HRTEM) images of (f) TCN50-C sample, (g,h) TCN50-D sample.
Figure 5. SEM images of (a) TiO2-C, (b) TiO2-D, (c) g-C3N4, (d) TCN50-D and (e) EDS elemental mappings of TCN50-D samples; high-resolution transmission electron microscopy (HRTEM) images of (f) TCN50-C sample, (g,h) TCN50-D sample.
Nanomaterials 10 00805 g005
Figure 6. (a) Photocatalytic degradation of RhB under visible light irradiation. (b) The photodegradation stability of RhB over TCN50-D sample. (c) Photocatalytic H2 evolution rates of TiO2, g-C3N4, TCN50-C and TCN50-D under visible light irradiation. (d) Photocatalytic mechanism for the charge transfer between g-C3N4 and TiO2 under visible light irradiation.
Figure 6. (a) Photocatalytic degradation of RhB under visible light irradiation. (b) The photodegradation stability of RhB over TCN50-D sample. (c) Photocatalytic H2 evolution rates of TiO2, g-C3N4, TCN50-C and TCN50-D under visible light irradiation. (d) Photocatalytic mechanism for the charge transfer between g-C3N4 and TiO2 under visible light irradiation.
Nanomaterials 10 00805 g006
Table 1. Surface and structural characterization of TiO2, g-C3N4 and TCNX composites.
Table 1. Surface and structural characterization of TiO2, g-C3N4 and TCNX composites.
SampleSBET (m2/g)Pore Volume (cm3/g)Average Pore Radius (nm)Crystallite Size (nm)
TiO2-D64.56490.15662.9914.3
TiO2-C28.18760.10236.7917.8
TCN50-D72.84730.14512.8612.3
TCN50-C29.27350.11176.0214.4
g-C3N47.13130.060411.34-

Share and Cite

MDPI and ACS Style

Zhang, B.; Peng, X.; Wang, Z. Noble Metal-Free TiO2-Coated Carbon Nitride Layers for Enhanced Visible Light-Driven Photocatalysis. Nanomaterials 2020, 10, 805. https://doi.org/10.3390/nano10040805

AMA Style

Zhang B, Peng X, Wang Z. Noble Metal-Free TiO2-Coated Carbon Nitride Layers for Enhanced Visible Light-Driven Photocatalysis. Nanomaterials. 2020; 10(4):805. https://doi.org/10.3390/nano10040805

Chicago/Turabian Style

Zhang, Bo, Xiangfeng Peng, and Zhao Wang. 2020. "Noble Metal-Free TiO2-Coated Carbon Nitride Layers for Enhanced Visible Light-Driven Photocatalysis" Nanomaterials 10, no. 4: 805. https://doi.org/10.3390/nano10040805

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