Next Article in Journal
Electrochemical Sensors Based on Conducting Polymers for the Aqueous Detection of Biologically Relevant Molecules
Next Article in Special Issue
Suitability of Different Titanium Dioxide Nanotube Morphologies for Photocatalytic Water Treatment
Previous Article in Journal
Assessment of the Influence of Crystalline Form on Cyto-Genotoxic and Inflammatory Effects Induced by TiO2 Nanoparticles on Human Bronchial and Alveolar Cells
Previous Article in Special Issue
Microwave-Assisted Synthesis of Chalcopyrite/Silver Phosphate Composites with Enhanced Degradation of Rhodamine B under Photo-Fenton Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

TiO2 Nanowires with Doped g-C3N4 Nanoparticles for Enhanced H2 Production and Photodegradation of Pollutants

1
College of Environment and Chemical Engineering, Nanchang Hangkong University, Nanchang 330063, China
2
Shandong Key Laboratory of Chemical Energy Storage and Novel Cell Technology, Liaocheng University, Liaocheng 252059, China
3
School of Ocean Science and Biochemistry Engineering, Fuqing Branch of Fujian Normal University, Fuqing 350300, China
4
School of Civil Engineering and Architecture, East China Jiaotong University, Nanchang 330013, China
*
Authors to whom correspondence should be addressed.
Submission received: 30 November 2020 / Revised: 10 January 2021 / Accepted: 18 January 2021 / Published: 19 January 2021
(This article belongs to the Special Issue Application of Nanomaterials in Photocatalysis)

Abstract

:
With the rapid consumption of fossil fuels, along with the ever-increasing environmental pollution, it is becoming a top priority to explore efficient photocatalysts for the production of renewable hydrogen and degradation of pollutants. Here, we fabricated a composite of g-C3N4/TiO2 via an in situ growth method under the conditions of high-temperature calcination. In this method, TiO2 nanowires with a large specific surface area could provide enough space for loading more g-C3N4 nanoparticles to obtain C3N4/TiO2 composites. Of note, the g-C3N4/TiO2 composite could effectively photocatalyze both the degradation of several pollutants and production of hydrogen, both of which are essential for environmental governance. Combining multiple characterizations and experiments, we found that the heterojunction constructed by the TiO2 and g-C3N4 could increase the photocatalytic ability of materials by prompting the separation of photogenerated carriers. Furthermore, the photocatalytic mechanism of the g-C3N4/TiO2 composite was also clarified in detail.

1. Introduction

The vigorous development of various industries drives rapid economic progress but causes excessive energy consumption and environmental pollution [1]. Such problems should not be underestimated; thus, finding efficient technology and sustainable energy has become the top priority. In recent years, hydrogen energy (H2) has been greatly popular with scientists because of its peculiarities of being renewable and clean. Considerable efforts have been devoted to the development of advanced technologies for harvesting hydrogen, and water splitting catalyzed by semiconductors is widely recognized as a promising approach to producing H2 [2,3].
TiO2 is a robust photocatalyst for water splitting for generating hydrogen, photodegradation, dye-sensitized solar cell biosensors, and other fields due to its high photostability, outstanding chemical stability, nontoxicity, and high efficiency [4,5]. However, the applications of TiO2 are still greatly limited by its intrinsic shortcomings. For example, the wide band gap (3.2 eV) and single crystalline phase can reduce the migration efficiency and lead to a high recombination rate for charges or limit the utilization of solar energy [6,7,8]. Therefore, numerous scientists have devoted themselves to solving these difficulties from different perspectives, such as the auto-doping of TiO2 [9], noble metal deposition, heterojunction construction, and ion doping [10,11,12,13,14,15,16,17]. Among these methods, the construction of heterojunctions is believed to be one favorable way [18].
Over the past decade, g-C3N4 has drawn worldwide attention due to its advantages of high stability, being green, being cheap, and being easy to synthesize, as well as unique electronic structure [19,20]. Furthermore, the morphology of materials also impacts the performance. The surface area of nanowires is dozens of times bigger than that of particles and has a strong absorption capacity. Meanwhile, nanowires can be filled with other nanomaterials to generate nanocomposites for greatly enhancing photocatalytic ability [21].
Herein, we propose a strategy of combining TiO2 nanowires and g-C3N4 nanoparticles for generating a nanocomposite, aiming at improving photocatalytic performance. Considering that the synthetic process should be environmentally friendly and low cost as well as efficient, we used commercial P25 powder and sodium hydroxide (NaOH) as raw materials to prepare TiO2 nanowires. Under the reaction conditions of high temperature, high pressure, and strong alkalinity, P25 reacted with NaOH to produce titanium, which convolved to form a tubular structure after the Na+ was exchanged in the washing process. Simultaneously, hydroxide was dissolved in the water, and oxide precipitated due to the different solubility. At last, g-C3N4 nanoparticles were loaded on the TiO2 nanowires to construct a novel g-C3N4/TiO2 heterojunction via an in situ growth method at a high calcination temperature. Furthermore, the photocatalytic ability of all the materials was examined by the photodegradation of pollutants and photocatalytic hydrogen production.

2. Experimental Section

2.1. Experimental Reagents

Titanium dioxide (TiO2, P25, nanoscale) and sodium hydroxide (NaOH, AR) were both obtained from Xilong Chemical Co., Ltd (Shantou China). Melamine (C3N3(NH2)3) was purchased from Sinopharm Chemical Reagent Co., Ltd (Shanghai, China).

2.2. Synthesis of TiO2 Nanowires

The TiO2 nanowires were prepared according to a previous report with some modifications [22]. Firstly, P25 (0.5 g) was dispersed into a 75 mL NaOH solution (10 mg/L) using an ultrasonic instrument for 30 min, and then was stirred for another 30 min. After that, the well-mixed reactant was poured into a Teflon reactor (100 mL) and reacted at 130 °C for 24 h. After cooling down to room temperature, the precursor was obtained through centrifugation and washed until the pH was 9. Afterwards, the above product was washed and stirred in 100 mL of HNO3 (0.1 mol/L) for 5 h. After pickling, the solution was purified with deionized water until the pH was 7 and centrifuged to obtain precipitates. The precipitate was dried in a vacuum at 60 °C for 24 h and then calcined at 600 °C for 3 h with a heating rate of 5 °C/min to obtain TiO2 nanowires.

2.3. The Preparation of g-C3N4/TiO2 Samples

g-C3N4 was prepared by calcining the melamine at 500 °C. The as-prepared TiO2 nanowire powder and 1 g of melamine were evenly ground, and then calcined at 540 °C for 4 h with a heating rate of 3 °C/min. The different mass ratios of g-C3N4/TiO2 composite (such as 20%, 30%, 40%, and 50%) were synthesized by changing the quality of g-C3N4 without the other conditions being changed.

2.4. Characterization

The phase and textual properties of the samples were tested by X-ray powder diffraction (XRD). The micromorphology and lattice structure were further analyzed with a scanning electron microscope (SEM) and transmission electron microscope (TEM). The chemical compositions and states were determined by X-ray photoelectron spectroscopy (XPS). Besides, the surface area and porosity were determined on a specific surface area tester. The UV-vis diffuse reflection spectra (UV-vis DRSs) of the samples were acquired with an ultraviolet and visible spectrophotometer with an integrating sphere; BaSO4 was used as the background material. The photocurrent tests were carried out with an electrochemical workstation (CHI 660D) with a standard three-electrode system.

2.5. The Photocatalytic Ability Tests

2.5.1. The Experiments for Photodegradation

The photodegradation tests of methyl orange and rose red were carried out under a 300 W xenon light with a light filter (λ > 420 nm). Firstly, 50 mg of as-prepared g-C3N4/TiO2 complex was ultrasonically dispersed in a beaker containing the pollutant (50 mL, 10 ppm) and then magnetically stirred in the dark for 60 min to achieve the absorption–desorption equilibrium, labeled C0. Afterwards, the treated solution was irradiated under the xenon light for 120 min to perform the photodegrade reaction. Then, 3 mL of the pollutant suspension was extracted at the given interval times and then centrifuged and labeled Ct. At last, the change in the concentration of the contaminant was measured with an ultraviolet–visible spectrophotometer.

2.5.2. Photocatalytic H2 Generation

The photocatalytic hydrogen evolution reaction was performed in a quartz glass at 279.15 K. Meanwhile, the quartz glass was connected with a closed circulation system, and a 300W xenon light served as the light source. Typically, g-C3N4/TiO2 complex (50 mg) was dispersed in the solution, which consisted of methyl alcohol (8 mL) and deionized water (72 mL); then, Ar was pumped into the system to remove the air. Afterwards, the above reactor was illuminated under a 300 W xenon light for 3 h. At intervals of 0.5 h, the generated hydrogen gas was quantified with a gas chromatograph (GC7900).

2.5.3. The Tests for Photocurrent

The photocurrent experiments for the g-C3N4/TiO2 composite were monitored with an electrochemical station (CHI660C) with 0.5 mol/L NaSO4 solutions as the electrolyte solution. For this, the Ag/AgCl electrode was used as the reference electrode, and the graphite electrode and prepared samples served as the auxiliary electrode and the working electrode, respectively. Typically, the working electrode could be fabricated by the following method: 2 mg samples were evenly dispersed in 1 mL of deionized water. Then, 100 μL of suspension liquid was coated uniformly on tin oxide conductive glass, dried at room temperature, and set aside.

3. Results and Discussion

3.1. Micromorphology and Lattice Structural Characteristics

The micromorphology of the samples was preliminarily studied through the SEM images (Figure 1). Obviously, the shape of TiO2 was wirelike with a smooth surface and uniform thickness (Figure 1a,b). Meanwhile, it can be seen that the length of the TiO2 nanowires was inhomogeneous. This was attributed to the fact that P25 reacted with NaOH, forming titanate, which was convolved to a linear structure after sodium ions were exchanged in the washing process. Besides, the reason for the different lengths may also have been the high-temperature calcination. As shown in Figure 1c, we can clearly observe that the g-C3N4 particles stuck tightly on the TiO2 nanowires. The longer TiO2 nanowires were able to provide sufficient supporting space for the in situ growth of g-C3N4, which made it possible to load g-C3N4 on the TiO2.
In addition, we used a transmission electron microscope (TEM) to further view the morphology, particle size, and lattice structure of the samples. Similarly, the morphology (Figure 2a,b) of the TiO2 and g-C3N4 was nanowires and particles, respectively. These was consistent with the SEM images. The lattice fringes at a distance of 0.352 nm (Figure 2c,d) corresponded to the (101) plane of TiO2 [22,23,24], while the lattice fringes at 0.326 nm (Figure 2d) could be assigned to the crystal plane of g-C3N4 [25,26]. Therefore, it could be identified that the g-C3N4/TiO2 was successfully synthesized. In addition, the crossed and overlapped lattice fringes of the partial g-C3N4 and TiO2 reflected the idea that the coupling of g-C3N4 and TiO2 was not simple surface contact but the presence of chemical bonding. During the reaction process, Ti(OH)4 particles of the TiO2 nanowires were in contact with the triazine structure in the melamine, and then, the precursor was polymerized with high-temperature calcination. Therefore, TiO2 nanowires closely contacted g-C3N4 via chemical bonds, thereby generating a heterojunction.

3.2. Crystal Phase and Textural Characteristics

The crystal form and crystallinity of the samples were analyzed by obtaining the X-ray diffraction (XRD) patterns. The XRD patterns of the materials displayed five strong diffraction peaks located at 25.2°, 37.7°, 47.7°, 53.9°, and 54.5° (Figure 3), which corresponded to the (101), (004), (200), (105), and (211) crystal planes of TiO2 (ICPDS card No.21-1272), respectively [27]. This result indicates the as-prepared TiO2 was anatase phase. For g-C3N4, the characteristic peaks at 13.1° and 27.4° were consistent with the standard card for g-C3N4 [28]. However, no distinct characteristic peak of g-C3N4 was observed on the XRD patterns of the composite, probably due to the low content of g-C3N4 [29].
In general, the pore characters and specific surface areas of materials have great influences on the photocatalytic performance of the materials. The N2 absorption–desorption isotherms and pore diameter distribution are shown in Figure 4. As shown in Figure 4a,b, the lack of hysteresis loops and lack of obvious pore size distribution peaks in the TiO2 curve indicate that there was no large pore structure or mesoporous structure [30]. With an increase in g-C3N4 loading content, we found that curves typical of 20% g-C3N4/TiO2, 30% g-C3N4/TiO2, 40% g-C3N4/TiO2, and 50% g-C3N4/TiO2 belonged to the type-IV absorption isotherm, with a H3 hysteresis loop at higher pressures (0.8–1.0), which clarified the presence of mesoporous structure in the samples [31,32]. The formation of the pore was on account of the accumulation of g-C3N4 nanoparticles on the surface of the TiO2 nanowires, and the g-C3N4 changed the surface roughness of the nanowires’ structure. As shown in Table 1, with the loading of g-C3N4 increasing from 20% to 50%, the specific surface areas of the samples firstly increased and then decreased, but the pore volume almost remained unchanged. This is because the excessive loading of g-C3N4 led to the aggregation of the g-C3N4 nanoparticles and thus reduced the specific surface area of the TiO2 nanowires [20].

3.3. Chemical State and Band Gap Analysis

Typically, XPS spectra are a powerful method for investigating the electronic structures of different elements in nanocomposites. The characteristic peaks of Ti, O, and C were detected in the survey of the g-C3N4/TiO2 composite (Figure 5a), which suggested the successful synthesis of the nanocomposite containing g-C3N4 and TiO2. The XPS spectra of C 1s (Figure 5b) displayed three peaks located at 284.6, 286.27, and 288.74 eV, corresponding to C-C, C-OH, and C=O (and COO) bonding, respectively [20,33]. In the case of O 1s (Figure 5c), the binding energies of 529.72, 530.72, and 532.05 eV were attributed to Ti-O, H-O, and C-O, respectively [34]. The XPS spectra of Ti 2p displayed two peaks for Ti4+ (Figure 5d), indicating that the Ti species were in the form of TiO2 [27]. After peak splitting fitting, the broad N 1s peak was divided into three contributions (Figure 5e), namely, C=N-C (398.71 eV), N-C3 (399.70 eV), and C-N-H (400.54 eV) [19].
The optical performance of the photocatalysts is shown in Figure 6a. Obviously, the absorption edge of the 40% g-C3N4/TiO2 composite showed a red shift compared with pure TiO2, which would be of benefit for enhancing the utilization of solar energy for improving the photocatalytic ability of materials. The band gap value (Figure 6b) of materials can be acquired by plotting (αhν)1/2 against hv and then extending the tangent line to intersect the coordinate axis [29]. The band gap value of 40% g-C3N4/TiO2 was lower than that of TiO2 and other materials (Figure S3 in Supporting information), which was of benefit for the photocatalytic reaction.

3.4. Photocatalytic Ability Analysis

During the experimental process, rose red and methyl orange were regarded as the target pollutants for evaluating the photodegradation performance of materials under ultraviolet visible light. As shown in Figure 7a, g-C3N4 is almost inactive for the photocatalytic degradation of methyl orange. Although the TiO2 is active for the degradation of methyl orange, the performance is very poor, with a 70% removal efficiency within 120 min of irradiation time. Impressively, the optimal proportional complex (40% g-C3N4/TiO2) can completely photodegrade the methyl orange in 60 min. Moreover, the efficiency of methyl orange removal by 40% g-C3N4/TiO2 in 20 min was as high as that by TiO2 in 120 min. Thus, it can be deduced that the heterojunction in the composite has a positive effect on the photocatalytic performance. Furthermore, the loading content of g-C3N4 also played a key role in the photodegradation performance. Excessive g-C3N4 aggregation led to a decrease in the specific surface area and thus limited the reaction. Therefore, we could further deduce that the heterojunction in the complex could inhibit the recombination of carriers facilitating the photocatalytic ability of materials [32].
To further investigate the excellent photodegradation ability of 40% g-C3N4/TiO2 complex, the photodegradation of rose red and methyl orange cycling experiments were also carried out. As shown in Figure 7b, it is obvious that 40% g-C3N4/TiO2 also had a favorable ability to treat rose red. Furthermore, the efficiency of methyl orange degradation by 40% g-C3N4/TiO2 was almost unchanged after five cycling experiments (Figure 7c), indicating that the 40% g-C3N4/TiO2 possessed high stability and reusability. In a word, the above experiments demonstrate that the as-prepared 40%TiO2/g-C3N4 composite has a bright future in applications for wastewater treatment.
The photodegradation principally made use of the photocatalytic oxidation performance of the photocatalysts, whereas the photocatalytic reduction could also play a part in plenty of fields. Therefore, the photocatalytic hydrogen generation experiments were conducted under ultraviolet–visible irradiation, and the hydrogen produced was quantified every 0.5 h. The H2 production performance of pure g-C3N4 and TiO2 was not satisfactory (Figure 8a,b). After coupling g-C3N4 with TiO2, the quantity and rate of hydrogen generation greatly increased, which can be explained as the presence of a heterojunction in the composite. It effectively prompted the separation of photogenerated carriers, thereby improving the performance of hydrogen evolution. However, when the content of g-C3N4 was 50%, the generation rate for H2 became slow because the excessive loading could block the active sites on the TiO2.

3.5. Photocurrent Analysis

Usually, the photocurrent response can reflect the transfer and separation of photogenerated charges under irradiation [35]. As depicted in Figure 9, the photocurrent signal of TiO2 was the weakest, indicating that the recombination of photogenerated charges was serious. However, the photocurrent response increased after g-C3N4 was loaded on the surface of the TiO2. The reason for this phenomenon was that the heterojunction prompted the separation of photogenerated electrons and holes, thereby providing more electrons that participated in hydrogen evolution [20].

3.6. Mechanism of Photocatalysis

Figure 10 shows the mechanism of the photodegradation process (a) and the hydrogen production (b). Under irradiation, g-C3N4 and TiO2 were excited to produce a mass of electrons and holes. According to a previous report [36], the conduction band value of g-C3N4 (−1.3eV vs. NHE) is more negative than that of pure TiO2 (−0.5 eV vs. NHE), while the valence band value of TiO2 (2.7eV vs. NHE) is more positive than that of g-C3N4 particles (1.4eV vs. NHE). Therefore, the partial electrons on the conduction band of g-C3N4 transfer to the conduction band of TiO2, enabling O2 to change into ·O2−. At the same time, the holes in the valence band of TiO2 were transferred to g-C3N4. The OH- and water were oxidated to hydroxyl radicals. At last, the methyl orange and rose red reacted with these radicals to form CO2 and H2O. Thereby, the electrons and holes were effectively separated. However, the electrons played a major role in the process of hydrogen production. The potential of H2 evolution (0 eV vs. NHE) was more positive than the conduction band value of TiO2 (−0.5 eV vs. NHE); thus, H+ could capture the electrons and change into H2.

4. Conclusions

In this work, TiO2 nanowires were firstly fabricated via the simple hydrothermal method, and then, g-C3N4 was prepared via thermal polymerization using melamine as the precursor. A novel g-C3N4/TiO2 heterojunction composite was successfully prepared by adjusting the mass ratio of TiO2 and melamine in the calcination process. After g-C3N4 was loaded on the surface of TiO2, the absorption range for light, the photocurrent response, and the specific surface area increased significantly, which were benefited by the presence of the heterojunction. The heterojunction in the composite could greatly prompt the separation of the photogenerated electrons and holes, thus enhancing the photocatalytic ability of the g-C3N4/TiO2. More importantly, the g-C3N4/TiO2 composite has the ability to both photocatalyze the degradation of different pollutants and produce hydrogen, which means it has broad application prospects. Meanwhile, the as-prepared sample also kept favorable stability after the cycling experiments. Therefore, this study provides a friendly and effective method for treating complex environmental pollutants and energy consumption.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2079-4991/11/1/254/s1. Figure S1: The SEM image of TiO2 nanowires, Figure S2: The EDX image of g-C3N4/TiO2 composite. Figure S3: The UV-vis diffuse reflectance spectra (a) and band gaps (b) of 20%g-C3N4/TiO2, 30%g-C3N4/TiO2 and 50% g-C3N4/TiO2 composite.

Author Contributions

Writing—original draft preparation, L.J. and F.Z.; the preparation of the catalysts, R.Z.; writing—review and editing, Y.X. and J.Z.; testing the photocatalytic performance of the catalysts, Y.L. and J.W.; validation, Y.H.; electrochemical measurements and analysis, R.G. and S.L.; formal analysis, Y.H.; Investigaion, H.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 21667019, 22066017, and 22002057); the Key Project of the Natural Science Foundation of Jiangxi Province (No. 20171ACB20016); the Jiangxi Province Major Academic and Technical Leaders Cultivating Object Program (No. 20172BCB22014); the Science and Technology Department of Jiangxi Province (No. 20181BCB18003 and 20181ACG70025); the Key Laboratory of Photochemical Conversion and Optoelectronic Materials, TIPC, CSA (No. PCOM201906); the Key Project of Science and Technology Research of the Jiangxi Provincial Department of Education (No. DA201602063 and GJJ191044); the Aviation Science Foundation of China (No. 2017ZF56020); the Fujian Key Laboratory of Measurement and Control System for Shore Environment (No. S1-KF1703); and the Doctor’s Start-Up Fund of Nanchang Hangkong University (EA201902286).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, Y.; Xie, Y.; Ling, Y.; Jiao, J.; Li, X.; Zhao, J. Facile construction of a molybdenum disulphide/zinc oxide nanosheet hybrid for an advanced photocatalyst. J. Alloy. Compd. 2019, 778, 761–767. [Google Scholar] [CrossRef]
  2. Xu, D.; Hai, Y.; Zhang, X.; Zhang, S.; He, R. Bi2O3 cocatalyst improving photocatalytic hydrogen evolution performance of TiO2. Appl. Surf. Sci. 2017, 400, 530–536. [Google Scholar] [CrossRef]
  3. Liu, Y.; Xu, C.; Xie, Y.; Yang, L.; Ling, Y.; Chen, L. Au–Cu nanoalloy/TiO2/MoS2 ternary hybrid with enhanced photocatalytic hydrogen production. J. Alloy. Compd. 2020, 820, 153440. [Google Scholar] [CrossRef]
  4. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, X.; Li, Z.; Shi, J.; Yu, Y. One-dimensional titanium dioxide nanomaterials: Nanowires, nanorods, and nanobelts. Chem. Rev. 2014, 114, 9346–9384. [Google Scholar] [CrossRef] [PubMed]
  6. 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, 16883–16890. [Google Scholar] [CrossRef] [PubMed]
  7. Chen, L.; Shen, L.; Nie, P.; Zhang, X.; Li, H. Facile hydrothermal synthesis of single crystalline TiOF2 nanocubes and their phase transitions to TiO2 hollow nanocages as anode materials for lithium-ion battery. Electrochim. Acta 2012, 62, 408–415. [Google Scholar] [CrossRef]
  8. Wen, C.Z.; Hu, Q.H.; Guo, Y.N.; Gong, X.Q.; Qiao, S.Z.; Yang, H.G. From titanium oxydifluoride (TiOF2) to titania (TiO2): Phase transition and non-metal doping with enhanced photocatalytic hy-drogen (H2) evolution properties. Chem. Commun. 2011, 47, 6138–6140. [Google Scholar] [CrossRef] [Green Version]
  9. Huang, Z.; Sun, Q.; Lv, K.; Zhang, Z.; Li, M.; Li, B. Effect of contact interface between TiO2 and g-C3N4 on the photoreactivity of g-C3N4/TiO2 photocatalyst: (001) vs. (101) facets of TiO2. Appl. Catal. B Environ. 2015, 164, 420–427. [Google Scholar] [CrossRef]
  10. Cao, S.-W.; Yuan, Y.-P.; Fang, J.; Shahjamali, M.M.; Boey, F.Y.C.; Barber, J.; Loo, S.J.C.; Xue, C. In-situ growth of CdS quantum dots on g-C3N4 nanosheets for highly efficient photocatalytic hydrogen generation under visible light irradiation. Int. J. Hydrog. Energy 2013, 38, 1258–1266. [Google Scholar] [CrossRef]
  11. 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]
  12. Jiang, F.; Yan, T.; Chen, H.; Sun, A.; Xu, C.; Wang, X. A g-C3N4–CdS composite catalyst with high visible-light-driven catalytic activity and photostability for methylene blue degradation. Appl. Surf. Sci. 2014, 295, 164–172. [Google Scholar] [CrossRef]
  13. Obregón, S.; Colón, G. Improved H2 production of Pt-TiO2/g-C3N4-MnOx composites by an efficient handling of photogenerated charge pairs. Appl. Catal. B Environ. 2014, 144, 775–782. [Google Scholar] [CrossRef]
  14. Wang, R.; Gu, L.; Zhou, J.; Liu, X.; Teng, F.; Li, C.; Shen, Y.; Yuan, Y. Quasi-polymeric metal–organic framework UiO-66/g-C3N4 heterojunctions for enhanced photocatalytic hydrogen evolution under visible light irradiation. Adv. Mater. Interfaces 2015, 2, 1500037. [Google Scholar] [CrossRef]
  15. Wang, Y.; Shi, R.; Lin, J.; Zhu, Y. Enhancement of photocurrent and photocatalytic activity of ZnO hybridized with graphite-like C3N4. Energy Environ. Sci. 2011, 4, 2922–2929. [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. Yuan, Y.-P.; Yin, L.-S.; Cao, S.-W.; Xu, G.-S.; Li, C.-H.; Xue, C. Improving photocatalytic hydrogen production of metal–organic framework UiO-66 octahedrons by dye-sensitization. Appl. Catal. B Environ. 2015, 168, 572–576. [Google Scholar] [CrossRef]
  18. Wang, X.-J.; Yang, W.-Y.; Li, F.-T.; Zhao, J.; Liu, R.-H.; Liu, S.-J.; Li, B. Construction of amorphous TiO2/BiOBr heterojunctions via facets coupling for enhanced photocatalytic activity. J. Hazard. Mater. 2015, 292, 126–136. [Google Scholar] [CrossRef]
  19. Hou, Y.; Yang, J.; Lei, C.; Yang, B.; Li, Z.; Xie, Y.; Zhang, X.; Lei, L.; Chen, J. Nitrogen vacancy structure driven photoeletrocatalytic degradation of 4-chlorophenol using porous graphitic carbon nitride nanosheets. ACS Sustain. Chem. Eng. 2018, 6, 6497–6506. [Google Scholar] [CrossRef]
  20. Zhou, Y.; Li, J.; Liu, C.; Huo, P.; Wang, H. Construction of 3D porous g-C3N4/AgBr/rGO composite for excellent visible light photocatalytic activity. Appl. Surf. Sci. 2018, 458, 586–596. [Google Scholar] [CrossRef]
  21. Truong, Q.D.; Le, T.S.; Hoa, T.H. Ultrathin TiO2 rutile nanowires enable reversible Mg-ion intercalation. Mater. Lett. 2019, 254, 357–360. [Google Scholar] [CrossRef]
  22. Yoshida, R.; Suzuki, Y.; Yoshikawa, S. Syntheses of TiO2(B) nanowires and TiO2 anatase nanowires by hydrothermal and post-heat treatments. J. Solid State Chem. 2005, 178, 2179–2185. [Google Scholar] [CrossRef]
  23. Yu, J.; Low, J.; Xiao, W.; Zhou, P.; Jaroniec, M. Enhanced photocatalytic CO2-reduction activity of anatase TiO2 by coexposed {001} and {101} facets. J. Am. Chem. Soc. 2014, 136, 8839–8842. [Google Scholar] [CrossRef] [PubMed]
  24. Cheng, H.; Selloni, A. Energetics and diffusion of intrinsic surface and subsurface defects on ana-tase TiO2(101). J. Chem. Phys. 2009, 131, 054703. [Google Scholar] [CrossRef]
  25. Tang, Y.; Zhang, Y.; Deng, J.; Wei, J.; Tam, H.L.; Chandran, B.K.; Dong, Z.; Chen, Z.; Chen, X. Nanotubes: Mechanical force-driven growth of elongated bending TiO2-based nanotubular materials for ultrafast rechargeable lithium ion batteries. Adv. Mater. 2014, 26, 6111–6118. [Google Scholar] [CrossRef]
  26. Li, Q.; Yue, B.; Iwai, H.; Kako, T.; Ye, J. Carbon nitride polymers sensitized with N-doped tantalic acid for visible light-induced photocatalytic hydrogen evolution. J. Phys. Chem. C 2010, 114, 4100–4105. [Google Scholar] [CrossRef]
  27. Sakata, Y.; Yoshimoto, K.; Kawaguchi, K.; Imamura, H.; Higashimoto, S. Preparation of a semi-conductive compound obtained by the pyrolysis of urea under N2 and the photocatalytic property under visible light irradiation. Catal. Today 2011, 161, 41–45. [Google Scholar] [CrossRef]
  28. Li, Y.; Fu, Y.; Zhu, M. Green synthesis of 3D tripyramid TiO2 architectures with assistance of aloe extracts for highly efficient photocatalytic degradation of antibiotic ciprofloxacin. Appl. Catal. B Environ. 2020, 260, 118149. [Google Scholar] [CrossRef]
  29. Li, K.; Huang, Z.; Zhu, S.; Luo, S.; Yan, L.; Dai, Y.; Guo, Y.; Yang, Y. Removal of Cr(VI) from water by a biochar-coupled g-C3N4 nanosheets composite and performance of a recycled photocatalyst in single and combined pollution systems. Appl. Catal. B Environ. 2019, 243, 386–396. [Google Scholar] [CrossRef]
  30. Yu, X.; Wu, P.; Qi, C.; Shi, J.; Feng, L.; Li, C.; Wang, L. Ternary-component reduced graphene ox-ide aerogel constructed by g-C3N4/BiOBr heterojunction and graphene oxide with enhanced photo-catalytic performance. J. Alloy. Compd. 2017, 729, 162–170. [Google Scholar] [CrossRef]
  31. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  32. Sun, C.; Xu, Q.; Xie, Y.; Ling, Y.; Hou, Y. Designed synthesis of anatase–TiO2 (B) biphase nanowire/ZnO nanoparticle heterojunction for enhanced photocatalysis. J. Mater. Chem. A 2018, 6, 8289–8298. [Google Scholar] [CrossRef]
  33. Hu, D.; Xie, Y.; Liu, L.; Zhou, P.; Zhao, J.; Xu, J.; Ling, Y. Constructing TiO2 nanoparticles patched nanorods heterostructure for efficient photodegradation of multiple organics and H2 production. Appl. Catal. B Environ. 2016, 188, 207–216. [Google Scholar] [CrossRef]
  34. Zou, H.; Song, M.; Yi, F.; Bian, L.; Liu, P.; Zhang, S. Simulated-sunlight-activated photocatalysis of Methyl Orange using carbon and lanthanum co-doped Bi2O3–TiO2 composite. J. Alloy. Compd. 2016, 680, 54–59. [Google Scholar] [CrossRef]
  35. Shao, P.; Tian, J.; Zhao, Z.; Shi, W.; Gao, S.; Cui, F. Amorphous TiO2 doped with carbon for visible light photodegradation of rhodamine B and 4-chlorophenol. Appl. Surf. Sci. 2015, 324, 35–43. [Google Scholar] [CrossRef] [Green Version]
  36. Pang, Y.; Li, Y.; Xu, G.; Hu, Y.; Kou, Z.; Feng, Q.; Lv, J.; Zhang, Y.; Wang, J.; Wu, Y. Z-scheme carbon-bridged Bi2O3/TiO2 nanotube arrays to boost photoelectrochemical detection performance. Appl. Catal. B Environ. 2019, 248, 255–263. [Google Scholar] [CrossRef]
Figure 1. SEM images of TiO2 (a,b) and g-C3N4/TiO2 composite (c).
Figure 1. SEM images of TiO2 (a,b) and g-C3N4/TiO2 composite (c).
Nanomaterials 11 00254 g001
Figure 2. TEM images of TiO2 (a) and g-C3N4/TiO2 (b); HRTEM images of TiO2 (c) and g-C3N4/TiO2 (d).
Figure 2. TEM images of TiO2 (a) and g-C3N4/TiO2 (b); HRTEM images of TiO2 (c) and g-C3N4/TiO2 (d).
Nanomaterials 11 00254 g002
Figure 3. The XRD patterns of TiO2, g-C3N4, and g-C3N4/TiO2 samples with different mass ratios.
Figure 3. The XRD patterns of TiO2, g-C3N4, and g-C3N4/TiO2 samples with different mass ratios.
Nanomaterials 11 00254 g003
Figure 4. The N2 adsorption–desorption isothermal curve (a); pore diameter distribution curves (b).
Figure 4. The N2 adsorption–desorption isothermal curve (a); pore diameter distribution curves (b).
Nanomaterials 11 00254 g004
Figure 5. The high-resolution XPS spectra of the 40% g-C3N4/TiO2 composite: survey (a), C 1s (b), O 1s (c), Ti 2p (d), and N 1s (e).
Figure 5. The high-resolution XPS spectra of the 40% g-C3N4/TiO2 composite: survey (a), C 1s (b), O 1s (c), Ti 2p (d), and N 1s (e).
Nanomaterials 11 00254 g005
Figure 6. The UV-vis diffuse reflectance spectra (a) and band gaps (b) of TiO2 and g-C3N4/TiO2 composite.
Figure 6. The UV-vis diffuse reflectance spectra (a) and band gaps (b) of TiO2 and g-C3N4/TiO2 composite.
Nanomaterials 11 00254 g006
Figure 7. The photodegradation curves for methyl orange (a) and the photodegradation curves for rose red (b) with g-C3N4, TiO2, and g-C3N4/TiO2 composite; the methyl orange cycling photodegradation experiments with 40% g-C3N4/TiO2 composite (c).
Figure 7. The photodegradation curves for methyl orange (a) and the photodegradation curves for rose red (b) with g-C3N4, TiO2, and g-C3N4/TiO2 composite; the methyl orange cycling photodegradation experiments with 40% g-C3N4/TiO2 composite (c).
Nanomaterials 11 00254 g007
Figure 8. Photocatalytic hydrogen production ability (a) and hydrogen production rate tests (b) of g-C3N4, TiO2, and g-C3N4/TiO2 samples.
Figure 8. Photocatalytic hydrogen production ability (a) and hydrogen production rate tests (b) of g-C3N4, TiO2, and g-C3N4/TiO2 samples.
Nanomaterials 11 00254 g008
Figure 9. Photocurrent tests of TiO2 and g-C3N4/TiO2 samples with different mass ratios.
Figure 9. Photocurrent tests of TiO2 and g-C3N4/TiO2 samples with different mass ratios.
Nanomaterials 11 00254 g009
Figure 10. Photocatalytic degradation by (a) and hydrogen production (b) mechanism of g-C3N4/TiO2 composite.
Figure 10. Photocatalytic degradation by (a) and hydrogen production (b) mechanism of g-C3N4/TiO2 composite.
Nanomaterials 11 00254 g010
Table 1. The specific surface area of samples.
Table 1. The specific surface area of samples.
SamplesSpecific Surface Area (m2/g)Pore Volume (cm3/g)
TiO211.9710.07
20% g-C3N4/TiO266.1810.53
30% g-C3N4/TiO280.6910.51
40% g-C3N4/TiO2131.1810.57
50% g-C3N4/TiO272.8710.48
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jiang, L.; Zeng, F.; Zhong, R.; Xie, Y.; Wang, J.; Ye, H.; Ling, Y.; Guo, R.; Zhao, J.; Li, S.; et al. TiO2 Nanowires with Doped g-C3N4 Nanoparticles for Enhanced H2 Production and Photodegradation of Pollutants. Nanomaterials 2021, 11, 254. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11010254

AMA Style

Jiang L, Zeng F, Zhong R, Xie Y, Wang J, Ye H, Ling Y, Guo R, Zhao J, Li S, et al. TiO2 Nanowires with Doped g-C3N4 Nanoparticles for Enhanced H2 Production and Photodegradation of Pollutants. Nanomaterials. 2021; 11(1):254. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11010254

Chicago/Turabian Style

Jiang, Liushan, Fanshan Zeng, Rong Zhong, Yu Xie, Jianli Wang, Hao Ye, Yun Ling, Ruobin Guo, Jinsheng Zhao, Shiqian Li, and et al. 2021. "TiO2 Nanowires with Doped g-C3N4 Nanoparticles for Enhanced H2 Production and Photodegradation of Pollutants" Nanomaterials 11, no. 1: 254. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11010254

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