Next Article in Journal
Recent Advances in Electrochemical Sensing of Hydrogen Peroxide (H2O2) Released from Cancer Cells
Next Article in Special Issue
Construction of Spindle-Shaped Ti3+ Self-Doped TiO2 Photocatalysts Using Triethanolamine-Aqueous as the Medium and Its Photoelectrochemical Properties
Previous Article in Journal
Microwave-Assisted Synthesis of N/TiO2 Nanoparticles for Photocatalysis under Different Irradiation Spectra
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Charge Transfer Channels in Reduced Graphene Oxide/Mesoporous TiO2 Nanotube Heterojunction Assemblies toward Optimized Photocatalytic Hydrogen Evolution

Shandong Provincial Key Laboratory of Molecular Engineering, School of Chemistry and Chemical Engineering, Qilu University of Technology (Shandong Academy of Sciences), Jinan 250353, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2022, 12(9), 1474; https://doi.org/10.3390/nano12091474
Submission received: 26 March 2022 / Revised: 18 April 2022 / Accepted: 23 April 2022 / Published: 26 April 2022
(This article belongs to the Special Issue Synthesis of TiO2 Nanoparticles and Their Catalytic Activity)

Abstract

:
Interface engineering is usually considered to be an efficient strategy to promote the separation and migration of photoexcited electron-hole pairs and improve photocatalytic performance. Herein, reduced graphene oxide/mesoporous titanium dioxide nanotube heterojunction assemblies (rGO/TiO2) are fabricated via a facile hydrothermal method. The rGO is anchored on the surface of TiO2 nanosheet assembled nanotubes in a tightly manner due to the laminated effect, in which the formed heterojunction interface becomes efficient charge transfer channels to boost the photocatalytic performance. The resultant rGO/TiO2 heterojunction assemblies extend the photoresponse to the visible light region and exhibit an excellent photocatalytic hydrogen production rate of 932.9 μmol h−1 g−1 under simulated sunlight (AM 1.5G), which is much higher than that of pristine TiO2 nanotubes (768.4 μmol h−1 g−1). The enhancement can be ascribed to the formation of a heterojunction assembly, establishing effective charge transfer channels and favoring spatial charge separation, the introduced rGO acting as an electron acceptor and the two-dimensional mesoporous nanosheets structure supplying a large surface area and adequate surface active sites. This heterojunction assembly will have potential applications in energy fields.

Graphical Abstract

1. Introduction

In recent years, the energy crisis and environmental pollution have become two major topics in the field of science and technology research [1,2]. The development and utilization of new technology has become the primary task of current research. Among various options, photocatalysis is favored for its low carbon footprint and use of renewable resources [3,4,5]. The most important thing is that it can directly use solar energy to generate new energy, such as hydrogen energy, and plays an important role in the degradation of pollutants and nitrogen fixation [6]. Therefore, it is considered to be an efficient strategy to solve the current increasingly serious environmental problems, and it has good application prospects.
In 1972, Honda and Fujishima [7] discovered that titanium dioxide (TiO2) could cause photocatalytic water-splitting. Since then, many scholars have begun to conduct a large number of experimental studies on TiO2 [8]. So far, TiO2 has obviously become the most widely studied photocatalyst due to its advantages of high stability and activity, super hydrophilicity, low cost and environmental friendliness in the field of photocatalysis [9,10]. However, from the perspective of the mechanism, it can be found that TiO2 has three fatal defects, which seriously limit its development. Firstly, TiO2 has a large band gap energy (3.2 eV), which means that the photogenerated electron-hole can only be activated by ultraviolet light with an energy higher than 3.2 eV (wavelength less than 387 nm), which only accounts for ~5% of sunlight. It was clear that visible light with a relatively long wavelength (almost 43% of sunlight) could not stimulate the photocatalytic activity of TiO2 [11,12]. Secondly, the excited photogenerated electrons and holes in TiO2 have a strong oxidability and reducibility, which leads to a rapid recombination and inefficient quantum efficiency. Studies have shown that most (about 90%) of the excited photogenerated electrons and holes of TiO2 have recombined before photocatalysis [13]. Thirdly, traditional TiO2 materials have a low specific surface area, which restricts the photocatalytic activity of TiO2 obviously [14,15].
In recent years, some methods have been adopted to improve the photocatalytic performance of TiO2. Firstly, the surface area of TiO2 could be improved by adjusting the morphology [16]. In addition, an excellent morphology can also improve the probability of charge transport by shortening the transmission path [17]. Secondly, TiO2 is modified by doping, metal deposition, and recombination to improve its activity. Doping can be used for improving the charge transfer efficiency and can be used for dragging down the bandgap [18]. Metal deposition can either form a Schottky barrier or an Ohmic contact, improving the efficiency of charge separation [19], and the introduction of metal can also broaden the absorption of light into the visible light region [20]. Recombination can improve charge separation efficiency through electron transfer [21]. Among various methods, heterojunction construction is a simple and practical method to effectively promote charge separation [22,23,24]. Graphene, as a representative of two-dimensional carbon materials, is composed of a single layer of carbon atoms with an sp2 hybrid orbital in the shape of a hexagonal honeycomb [25]. Graphene oxide (GO) is prepared by embedding O atoms into the c-scaffold of graphene to make sp2 and sp3 domains exist simultaneously in the structure, which promotes the expansion of multiple interactions. On this basis, reduced graphene oxide (rGO) can provide more interfacial polarization sites and absorption sites [26]. Thus, rGO has obvious application prospects in many fields including energy and environment, along with thermal and electrical properties [27,28,29]. In recent years, some reports had been reported on the application of rGO/TiO2 composite materials in the field of photocatalysis. Balsamo et al. successfully prepared a TiO2-rGO composite structure using the one-pot method by solar irradiation, which is simple and green. The degradation rate of 2,4-dichlorophenoxyacetic acid after irradiation for 3 h reached 97%. The enhanced absorption of the composite is due to the interaction between the TiO2 and rGO [30]. Zouzelka et al. optimized rGO/TiO2 composites by electrophoretic deposition. The presence of rGO can, to some extent, promote the photocatalytic degradation of 4-chlorophenol. Compared with TiO2, the photocatalytic degradation rate of the composite material is several times higher, and this enhanced performance is due to the charge transfer promoting the formation of hydroxyl radical [31]. Firstly, rGO could change the band gap width of TiO2 to expand its light response range. Secondly, it can be used as an electron transporter, which can effectively prevent photogenerated electron hole recombination and significantly prolong the lifetime of the electron-hole pair. In addition, a large specific surface area can provide more active sites for photocatalytic reactions [32]. Therefore, rGO/TiO2 composite materials maybe have good application prospects in the field of photocatalysis due to their high charge separation efficiency. Although these rGO/TiO2 composites indeed improved the photocatalytic performance obviously, they are still far from being practical applications. The question of how to further promote the charge separation of rGO/TiO2-based materials via interface engineering is still a great challenge.
Here, novel reduced graphene oxide/mesoporous titanium dioxide nanotube heterojunction assemblies (rGO/TiO2) are fabricated via a facile hydrothermal method by utilizing mesoporous TiO2 nanosheet-assembled nanotubes as the host. For one thing, the morphology of TiO2 was regulated into the nanotube structure assembled by the mesoporous nanosheet, which increased the specific surface area of the material and exposed more active sites. Besides, a tubular structure increased the multiple reflection of sunlight inside the tube and improved the utilization rate of light. For another thing, rGO, as an electron acceptor, was introduced to form rGO/TiO2 heterojunction assemblies, which contributed to improve the spatial charge separation efficiency and extend the photoresponse to the visible light region. The rGO/TiO2 heterojunction assemblies significantly improved the photocatalytic activity, and the photocatalytic hydrogen production rate was up to 932.9 umol h−1 g−1, 1.2 times higher than that of pristine TiO2. The novel rGO/TiO2 heterojunction assemblies may provide new insights for the fabrication of TiO2-based photocatalysts with a high performance.

2. Materials and Methods

2.1. Chemicals

Titanium oxysulfate (TiOSO4), potassium permanganate (KMnO4), sodium nitrate (NaNO3) and sulfuric acid (H2SO4) were purchased from Aladdin Chemical (Shanghai, China). Ethanol (EtOH), ether (C2H5OC2H5) and glycerol (C3H5(OH)3) were purchased from Tianjin Kermio Chemical (Tianjin, China). None of the chemicals were further purified.

2.2. Preparation of TiO2 Nanotube

The typical method for the synthesis of a TiO2 nanotube is based on the literature, and the specific steps are as follows [33]. 1 g TiOSO4 was added to 18 mL EtOH with intense stirring for 30 min, after which 9 mL C3H5(OH)3 and 9 mL C2H5OC2H5 were added drop by drop, stirred overnight, and then transferred to the Teflon reactor at 170 °C for 10 h. After the reaction, the obtained precursor was collected by repetitive centrifugation at 8000 rpm for 10 min and washed several times by deionized water and ethanol and dried at 60 °C for 12 h. After drying, the samples were transferred to a Muffle furnace and kept at 600 °C for 4 h. Finally, the TiO2 nanotube sample was obtained.

2.3. Preparation of rGO/TiO2

The formation process of rGO/TiO2 is shown in Figure 1a. rGO/TiO2 samples are synthesized on the basis of TiO2, and rGO is prepared by the hummer method using KMnO4, NaNO3 and H2SO4 as raw materials [34,35]. The prepared 5 mg rGO was added to 18 mL EtOH, stirred and ultrasonicated for 20 min, after which 1 g TiOSO4 was added. Then, 9 mL C3H5(OH)3 and 9 mL C2H5OC2H5 were added drop by drop, stirred overnight, and transferred to the Teflon-lined autoclave at 170 °C for 10 h. After the reaction, the resulting product was centrifuged and washed with deionized water and anhydrous ethanol 3 times and dried for later use. After drying, the samples were transferred to a Muffle furnace for calcination at 600 °C for 4 h. Finally, rGO/TiO2 samples were obtained. According to the feeding ratio, the content of rGO on TiO2 is 1 wt%.

3. Results and Discussion

3.1. Crystal Structure and Optical Absorption of rGO/TiO2

Figure 1b shows the X-ray diffraction (XRD) patterns of GO, TiO2 and GO/TiO2. The multiple diffraction peaks located at 25.3, 37.8, 48.0, 55.1, 62.8, 70.2 and 75.0° can be assigned to the (101), (004), (200), (211), (204), (116) and (220) planes of anatase TiO2 (JCPDS 21-1272), respectively [36]. This indicates the successful preparation of anatase TiO2. The characteristic peak positions of rGO (Figure 1b and Figure S1) are about 26° and 42.5°, which is consistent with the literature [37]. Figure 1b shows the XRD pattern of the rGO/TiO2 composite material, which is basically consistent with that of TiO2, indicating that the introduction of rGO does not destroy the crystal structure of TiO2. This also implies the stability of the material. However, the increased strength of the (101) crystal plane preliminarily indicates the successful introduction of rGO. The reason for this stems from the coincidence of the position of the TiO2 (101) crystal plane (25.3°) and rGO (26°), and no single peak of rGO can be observed. As shown in Figure 1c, UV-vis was applied to obtain the absorption property of the materials. TiO2 has a strong light absorption capacity when the wavelength is less than 400 nm, which is the inherent absorption property of crystal anatase TiO2. The absorption intensity of rGO/TiO2 was higher than that of TiO2, indicating the enhancement of the composites. The photoresponse region of the material extended the UV absorption to the visible light region, which was attributed to the modification of rGO. Figure 1d shows FTIR of TiO2 and rGO/TiO2, the principal vibration of TiO2 corresponding to the wide band in the range of 400−800 cm−1. The significant band at 670 cm−1 corresponds to the bending vibration mode of the Ti–O–Ti bond. For the rGO/TiO2 heterostructure, the wide absorption at low frequencies is similar to the spectrum of the TiO2 and can therefore be attributed to the vibration of the TiO2. In addition, the peak broadening of rGO/TiO2 relative to TiO2 below 1000 cm−1 may be due to the superposition of the peak existing in the Ti–O–C vibration (798 cm−1) on Ti–O–Ti. The emerging absorption band at around 1630 cm−1 can be associated with the bone vibration of the graphene sheet, indicating the successful loading of rGO [38]. The weak peak was due to the low amount of rGO on the TiO2 surface (about 1 wt%).

3.2. Morphology of rGO/TiO2 Assembly

A scanning electron microscope (SEM, Hitachi, S-4800, Tokyo, Japan) and transmission electron microscope (TEM, JEOL, JEM-2100, Tokyo, Japan) were used to observe the microstructure of TiO2 and rGO/TiO2. It can be clearly seen that TiO2 shows a very uniform tubular structure, assembled by many sheets, with a diameter of about 700 nm (Figure 2a,b). Compared with TiO2, the surface of rGO/TiO2 becomes slightly rough, and the diameter of the tube increases slightly (Figure 2c,d). The sheet structure can increase the specific surface area and expose more active sites. The tubular structure formed by the self-assembly of multiple nanosheets can further increase the specific surface area of the material. Secondly, the tubular structure can reflect and utilize the sunlight efficiently, which improves the utilization rate of light and is conducive to improving the photocatalytic activity. In addition, the heterostructure formed by TiO2 and rGO is beneficial to the electron transfer between materials, which reduces the electron-hole recombination rate and improves the charge separation efficiency and photocatalytic activity. TEM images of rGO/TiO2, as shown in Figure 2e, illustrate a nanosheet-assembled tubular structure, and the light-colored part on the surface is a layer of rGO, which is consistent with the results obtained by SEM. The rGO could be anchored on the surface of the nanosheets due to the laminated effect, thus forming efficient heterojunctions. The HRTEM images (Figure 2f,g) of rGO/TiO2 show that the edge of the sample is a layer of rGO and that the inner layer is TiO2. As shown in Figure 2g, clear lattice fringes can be seen with a lattice spacing of about 0.35 nm corresponding to the (101) crystal plane of anatase TiO2. Meanwhile, it can be clearly seen that the lattice spacing of rGO is 0.34 nm, which further indicates the successful preparation of the rGO/TiO2 heterojunction assembly material. Elemental mapping (Figure 2h–j) shows that O, C and Ti elements in rGO/TiO2 are evenly dispersed. The above information demonstrates the successful manufacturing of the rGO/TiO2 heterojunction assembly.

3.3. Photocatalytic Performance of rGO/TiO2

The photocatalytic performance of the catalysts for hydrogen production under favorable conditions of the catalyst Pt was conducted. As shown in Figure 3a, the hydrogen production performance of rGO is nearly 0, indicating that the presence of rGO does not contribute to hydrogen production, which is consistent with the reports in the literature. The hydrogen production quantity of rGO/TiO2 is 932.9 umol h−1 g−1, which is 1.2 times higher than that of original TiO2 (768.4 umol h−1 g−1). We can clearly see the excellent photocatalytic performance of rGO/TiO2, which is significantly better than in the previous literature (Table S1). This is mainly attributed to the formation of a heterojunction between TiO2 and rGO, which promotes the spatial charge separation and inhibits the recombination of electron-hole pairs. Secondly, the design of an rGO sheet supported on a TiO2 sheet is more conducive to external proton reduction. This 2D/2D design can also maximize the interface contact area, integrate the advantages of each 2D component, and facilitate the photocatalytic reaction. Finally, the nanosheet-assembled tubular structure provides a large specific surface area and sufficient surface active sites for the photocatalytic reaction. Under sunlight irradiation (AM 1.5G), the special microstructure can realize multiple reflections within the tubular structure, improving the utilization rate of sunlight. To investigate the stability of catalysts, cyclic tests were performed (Figure 3b). Five cycles of rGO/TiO2 were tested, and the hydrogen production rates were almost identical for each cycle. Besides, the amount of hydrogen production was not attenuated, indicating that rGO/TiO2 possessed a high stability, which was favorable for practical applications.
Figure 3c shows the electrochemical impedance spectra (EIS, Shanghai Chenhua Instrument Co., Ltd., CHI760E, Shanghai, China) of TiO2 and rGO/TiO2 under AM 1.5G light irradiation. It can be clearly seen that the arc radius of the EIS diagram of rGO/TiO2 is smaller than that of TiO2, indicating that rGO/TiO2 has a lower electrochemical impedance under light irradiation. This result clearly indicates that there is a positive synergistic effect between TiO2 and rGO on the enhancing electron migration, which has a more efficient charge carrier separation and faster electron transfer to reduce the electron-hole recombination rate and promote the photocatalytic hydrogen generation [39]. In order to analyze the semiconductor type and band structure of TiO2, the Mott–Schottky curve was tested for TiO2. As shown in Figure 3d, the slope of the curve is positive, which is typical for n-type semiconductors [40]. At the same time, the flat band potential of TiO2 (−1.05 eV) can be obtained from the slope of the curve, and the conduction band of TiO2 is −0.81 eV [41]. Combined with UV-vis, the valence band position is determined at 2.31 eV.

3.4. Photocatalytic Mechanism of rGO/TiO2

Based on the above, the mechanism diagram of photocatalytic hydrogen production of the rGO/TiO2 heterojunction was made, as shown in Scheme 1. When the rGO/TiO2 catalyst is irradiated by light, the electrons on TiO2 are excited to the conduction band (CB), leaving holes in the valence band (VB). Meanwhile, TiO2 will come into contact with rGO to form a built-in electric field. Since rGO is an electron acceptor, electrons will transfer from TiO2 CB to rGO. Additionally, the reduction reaction H+ converts to H2 on rGO. The improved photocatalytic performance is attributed to an electron derived from the rGO/TiO2 heterojunction, which reduces the recombination and improves the spatial charge separation efficiency. Such high photocatalytic properties are inseparable from the structure itself. The nanosheet-assembled hollow tubular structures provide a large specific surface area and adequate surface active sites for photocatalytic reactions. In addition, the nanotube structure can improve the utilization rate of light energy. The reason for this is that light can be reflected and used multiple times in the cavity. These factors all contribute to improving the photocatalytic hydrogen evolution of the rGO/TiO2 heterojunction assembly.

4. Conclusions

In summary, an rGO/TiO2 heterojunction assembly was prepared by a step hydrothermal method, and the H2 production rate was much faster than that of the original TiO2 nanotubes under AM 1.5G irradiation. This was mainly decided by the following two aspects. Firstly, the rGO heterostructure with TiO2 promoted the electron transfer, reduced the electron recombination rate, and improved the spatial charge efficiency. The design of rGO sheets supported by TiO2 sheets is more conducive to the reduction of external protons. Secondly, the advantage of the special structure was that the nanosheet-assembled hollow tube provided a large specific surface area. The combination of the two 2D structures maximizes the interface contact area and exposes enough reaction sites on the surface, which is favorable for photocatalytic hydrogen production. This novel strategy provides new ideas for the fabrication of other heterojunction photocatalysts with an efficient solar energy conversion.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12091474/s1. Characterizations, Photocatalytic activity, Photoelectrochemical measurements [42,43,44,45,46,47]. Figure S1: XRD patterns of rGO, Table S1: Comparison with other TiO2 and rGO composite catalysts.

Author Contributions

Conceptualization, Z.L., S.W. and W.Z.; methodology, D.Y.; software, L.G.; validation, T.C., Y.M. and X.H.; formal analysis, X.Z.; investigation, H.C.; resources, B.H.; data curation, S.Z., Y.W. and Y.G.; writing—original draft preparation, Z.L. and D.Y.; writing—review and editing, S.W. and W.Z.; visualization, W.Z.; supervision, S.W. and W.Z.; project administration, W.Z.; funding acquisition, W.Z. 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 (Grant Number: 21871078, 52172206); the Shandong Province Natural Science Foundation (Grant Number: ZR2021MB016), and the Development plan of Youth Innovation Team in Colleges and Universities of Shandong Province.

Data Availability Statement

Data presented in this article are available at request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Uddin, N.; Zhang, H.; Du, Y.; Jia, G.; Wang, S.; Yin, Z. Structural-phase catalytic redox reactions in energy and environmental applications. Adv. Mater. 2020, 32, 1905739. [Google Scholar] [CrossRef] [PubMed]
  2. Jeong, G.; Sasikala, S.; Yun, T.; Lee, G.; Lee, W.; Kim, S. Nanoscale assembly of 2d materials for energy and environmental applications. Adv. Mater. 2020, 32, 1907006. [Google Scholar] [CrossRef] [PubMed]
  3. Fang, B.; Xing, Z.; Sun, D.; Li, Z.; Zhou, W. Hollow semiconductor photocatalysts for solar energy conversion. Adv. Powder Mater. 2022, 1, 100021. [Google Scholar] [CrossRef]
  4. Wang, W.; Deng, C.; Xie, S.; Li, Y.; Zhang, W.; Sheng, H.; Chen, C.; Zhao, J. Photocatalytic c–c coupling from carbon dioxide reduction on copper oxide with mixed-valence copper(I)/copper(II). J. Am. Chem. Soc. 2021, 143, 2984–2993. [Google Scholar] [CrossRef] [PubMed]
  5. Jian, S.; Tian, Z.; Hu, J.; Zhang, K.; Zhang, L.; Duan, G.; Yang, W.; Jiang, S. Enhanced visible light photocatalytic efficiency of La-doped ZnO nanofibers via electrospinning-calcination technology. Adv. Powder Mater. 2022, 1, 100004. [Google Scholar] [CrossRef]
  6. Ma, B.; Blanco, M.; Calvillo, L.; Chen, L.; Chen, G.; Lau, T.-C.; Dražić, G.; Bonin, J.; Robert, M.; Granozzi, G. Hybridization of Molecular and Graphene Materials for CO2 Photocatalytic Reduction with Selectivity Control. J. Am. Chem. Soc. 2021, 143, 8414–8425. [Google Scholar] [CrossRef]
  7. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  8. Gong, H.; Wang, L.; Zhou, K.; Zhang, D.; Zhang, Y.; Adamaki, V.; Bowen, C.; Sergejevs, A. Improved photocatalytic performance of gradient reduced TiO2 ceramics with aligned pore channels. Adv. Powder Mater. 2021, 1, 100025. [Google Scholar] [CrossRef]
  9. Qiu, P.; Li, W.; Thokchom, B.; Park, B.; Cui, M.; Zhao, D.; Khim, J. Uniform core–shell structured magnetic mesoporous TiO2 nanospheres as a highly efficient and stable sonocatalyst for the degradation of bisphenol-A. J. Mater. Chem. A 2015, 3, 6492–6500. [Google Scholar] [CrossRef]
  10. Hao, Z.; Chen, Q.; Dai, W.; Ren, Y.; Zhou, Y.; Yang, J.; Xie, S.; Shen, Y.; Wu, J.; Chen, W.; et al. Oxygen-deficient blue TiO2 for ultrastable and fast lithium storage. Adv. Energy Mater. 2020, 10, 1903107. [Google Scholar] [CrossRef]
  11. Wang, H.; Hu, X.; Ma, Y.; Zhu, D.; Li, T.; Wang, J. Nitrate-Group-Grafting-Induced Assembly of Rutile TiO2 Nanobundles for Enhanced Photocatalytic Hydrogen Evolution. Chin. J. Catal. 2020, 41, 95–102. [Google Scholar] [CrossRef]
  12. Yu, L.; Zhang, G.; Liu, C.; Lan, H.; Liu, H.; Qu, J. Interface Stabilization of Undercoordinated Iron Centers on Manganese Oxides for Nature-inspired Peroxide Activation. ACS Catal. 2018, 8, 1090–1096. [Google Scholar] [CrossRef]
  13. Skinner, D.; Colombo, D.; Cavaleri, J.; Bowman, R. Femtosecond Investigation of Electron Trapping in Semiconductor Nanoclusters. J. Phys. Chem. 1995, 99, 7853–7856. [Google Scholar] [CrossRef]
  14. Lan, K.; Liu, Y.; Zhang, W.; Liu, Y.; Elzatahry, A.; Wang, R.; Xia, Y.; Al-Dhayan, D.; Zheng, N.; Zhao, D. Uniform Ordered Two-Dimensional Mesoporous TiO2 Nanosheets from Hydrothermal-Induced Solvent-Confined Monomicelle Assembly. J. Am. Chem. Soc. 2018, 140, 4135–4143. [Google Scholar] [CrossRef] [Green Version]
  15. Zhao, Y.; Zhao, Y.; Shi, R.; Wang, B.; Waterhouse, G.I.N.; Wu, L.; Tung, C.; Zhang, T. Tuning Oxygen Vacancies in Ultrathin TiO2 Nanosheets to Boost Photocatalytic Nitrogen Fixation up to 700 nm. Adv. Mater. 2019, 31, 1806482. [Google Scholar] [CrossRef]
  16. Wu, J.; Qiao, P.; Li, H.; Xu, Y.; Yang, W.; Yang, F.; Lin, K.; Pan, K.; Zhou, W. Engineering surface defects on two-dimensional ultrathin mesoporous anatase TiO2 nanosheets for efficient charge separation and exceptional solar-driven photocatalytic hydrogen evolution. J. Mater. Chem. C 2020, 8, 3476. [Google Scholar] [CrossRef]
  17. Bae, D.; Pedersen, T.; Seger, B.; Malizia, M.; Kuznetsov, A.; Hansen, O.; Chorkendorffa, I.; Vesborg, P. Back-illuminated Si photocathode: A combined experimental and theoretical study for photocatalytic hydrogen evolution. Energy Environ. Sci. 2015, 8, 650–660. [Google Scholar] [CrossRef] [Green Version]
  18. Yi, L.; Ci, S.; Luo, S.; Shao, P.; Hou, Y.; Wen, Z. Scalable and Low-cost Synthesis of Black Amorphous Al-Ti-O Nanostructure for High- efficient Photothermal Desalination. Nano Energy 2017, 41, 600–608. [Google Scholar] [CrossRef]
  19. Bae, D.; Kanellos, G.; Faasse, G.M.; Dražević, E.; Venugopal, A.; Smith, W.A. Design principles for efficient photoelectrodes in solar rechargeable redox flow cell applications. Commun. Mater. 2020, 1, 17. [Google Scholar] [CrossRef]
  20. Abed, J.; Rajput, N.S.; El Moutaouakil, A.; Jouiad, M. Recent Advances in the Design of Plasmonic Au/TiO2 Nanostructures for Enhanced Photocatalytic Water Splitting. Nanomaterials 2020, 10, 2260. [Google Scholar] [CrossRef]
  21. Song, W.; Ji, J.; Guo, K.; Wang, X.; Wei, X.; Cai, Y.; Tan, W.; Li, L.; Sun, J.; Tang, C.; et al. Solid-phase impregnation promotes Ce doping in TiO2 for boosted denitration of CeO2/TiO2 catalysts. Chin. Chem. Lett. 2022, 33, 935–938. [Google Scholar] [CrossRef]
  22. Zhu, Y.; Shah, M.; Wang, C. Insight into the role of Ti3+ in photocatalytic performance of shuriken-shaped BiVO4/TiO2−x heterojunction. Appl. Catal. B Environ. 2017, 203, 526–532. [Google Scholar] [CrossRef]
  23. Wang, H.; Liu, J.; Xiao, X.; Meng, H.; Wu, J.; Guo, C.; Zheng, M.; Wang, X.; Guo, S.; Jiang, B. Engineering of SnO2/TiO2 heterojunction compact interface with efficient charge transfer pathway for photocatalytic hydrogen evolution. Chin. Chem. Lett. 2022, in press. [Google Scholar] [CrossRef]
  24. Jia, M.; Yang, Z.; Xiong, W.; Cao, J.; Xiang, Y.; Peng, H.; Jing, Y.; Zhang, C.; Xu, H.; Song, P. Magnetic heterojunction of oxygen-deficient Ti3+-TiO2 and Ar-Fe2O3 derived from metal-organic frameworks for efficient peroxydisulfate (PDS) photo-activation. Appl. Catal. B Environ. 2021, 298, 120513. [Google Scholar] [CrossRef]
  25. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [Green Version]
  26. Sharma, D.; Menon, A.; Bose, S. Graphene templated growth of copper sulphide ‘flowers’ can suppress electromagnetic interference. Nanoscale Adv. 2020, 2, 3292–3303. [Google Scholar] [CrossRef]
  27. Wang, X.; Meng, G.; Zhu, C.; Huang, Z.; Qian, Y.; Sun, K.; Zhu, X. A generic synthetic approach to large-scale pristine-graphene/metal-nanoparticles hybrids. Adv. Funct. Mater. 2013, 23, 5685. [Google Scholar] [CrossRef]
  28. Qiu, B.; Xing, M.; Zhang, J. Mesoporous TiO2 nanocrystals grown in situ on graphene aerogels for high photocatalysis and lithium-ion batteries. J. Am. Chem. Soc. 2014, 136, 5852–5855. [Google Scholar] [CrossRef]
  29. Xu, C.; Wang, C.; He, X.; Lyu, M.; Wang, S.; Wang, L. Processable graphene oxideembedded titanate nanofiber membranes with improved filtration performance. J. Hazard. Mater. 2017, 325, 214–222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Balsamo, S.; Fiorenza, R.; Condorelli, M.; Pecoraro, R.; Brundo, M.; Presti, F.; Sciré, S. One-Pot Synthesis of TiO2-rGO Photocatalysts for the Degradation of Groundwater Pollutants. Materials 2021, 14, 5938. [Google Scholar] [CrossRef]
  31. Zouzelka, R.; Remzova, M.; Plšek, J.; Brabec, L.; Rathousky, J. Immobilized rGO/TiO2 Photocatalyst for Decontamination of Water. Catalysts 2019, 9, 708. [Google Scholar] [CrossRef] [Green Version]
  32. Wang, S.; Cai, J.; Mao, J.; Li, S.; Shen, J.; Gao, S.; Huang, J.; Wang, X.; Parkin, I.P.; Lai, Y. Defective black Ti3+ self-doped TiO2 and reduced graphene oxide composite nanoparticles for boosting visible-light driven photocatalytic and photoelectrochemical activity. Appl. Surf. Sci. 2019, 467–468, 45–55. [Google Scholar] [CrossRef]
  33. Zhang, X.; Hu, W.; Zhang, K.; Wang, J.; Sun, B.; Li, H.; Qiao, P.; Wang, L.; Zhou, W. Ti3+ Self-Doped Black TiO2 Nanotubes with Mesoporous Nanosheet Architecture as Efficient Solar-Driven Hydrogen Evolution Photocatalysts. ACS Sustain. Chem. Eng. 2017, 5, 6894–6901. [Google Scholar] [CrossRef]
  34. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved synthesis of graphene oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef]
  35. Cote, L.J.; Kim, F.; Huang, J. Langmuir-blodgett assembly of graphite oxide single layers. J. Am. Chem. Soc. 2009, 131, 1043–1049. [Google Scholar] [CrossRef]
  36. Yu, C.; Li, M.; Yang, D.; Pan, K.; Yang, F.; Xu, Y.; Yuan, L.; Qu, Y.; Zhou, W. NiO nanoparticles dotted TiO2 nanosheets assembled nanotubes P-N heterojunctions for efficient interface charge separation and photocatalytic hydrogen evolution. Appl. Surf. Sci. 2021, 568, 150981. [Google Scholar] [CrossRef]
  37. Chen, C.; Yang, S.; Ding, J.; Wang, G.; Zhong, L.; Zhao, S.; Zang, Y.; Jiang, J.; Ding, L.; Zhao, Y.; et al. Non-covalent self-assembly synthesis of AQ2S@rGO nanocomposite for the degradation of sulfadiazine under solar irradiation: The indispensable effect of chloride. Appl. Catal. B Environ. 2021, 298, 120495. [Google Scholar] [CrossRef]
  38. Sang, Y.; Zhao, Z.; Tian, J.; Hao, P.; Jiang, H.; Liu, H.; Claverie, J.P. Enhanced Photocatalytic Property of Reduced Graphene Oxide/TiO2 Nanobelt Surface Heterostructures Constructed by an In Situ Photochemical Reduction Method. Small 2014, 18, 3775–3782. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Fu, F.; Zhou, F.; Yang, X.; Zhang, D.; Chen, Y. Synergistic effect of RGO/TiO2 nanosheets with exposed (001) facets for boosting visible light photocatalytic activity. Appl. Surf. Sci. 2020, 510, 145451. [Google Scholar] [CrossRef]
  40. Kronik, L.; Shapira, Y. Surface photovoltage phenomena: Theory, experiment, and applications. Surf. Sci. Rep. 1999, 37, 1–206. [Google Scholar] [CrossRef]
  41. Yang, F.; Li, H.; Pan, K.; Wang, S.; Sun, H.; Xie, Y.; Xu, Y.; Wu, J.; Zhou, W. Engineering Surface N-Vacancy Defects of Ultrathin Mesoporous Carbon Nitride Nanosheets as Efficient Visible-Light-Driven Photocatalysts. Sol. RRL 2020, 5, 2000610. [Google Scholar] [CrossRef]
  42. Gupta, B.; Melvin, A.; Matthews, T.; Dhara, S.; Dash, S.; Tyagi, A. Facile gamma radiolytic methodology for TiO2-rGO synthesis: Effect on photo-catalytic H2 evolution. Int. J. Hydrogen Energy 2015, 40, 5815–5823. [Google Scholar] [CrossRef]
  43. Bharad, P.; Sivaranjani, K.; Gopinath, C. A rational approach towards enhancing solar water splitting: A case study of Au-RGO/N-RGO-TiO2. Nanoscale 2015, 7, 11206–11215. [Google Scholar] [CrossRef]
  44. Ma, J.; Dai, J.; Duan, Y.; Zhang, J.; Qiang, L.; Xue, J. Fabrication of PANI-TiO2/rGO hybrid composites for enhanced photocatalysis of pollutant removal and hydrogen production. Renew. Energ. 2020, 156, 1008–1018. [Google Scholar] [CrossRef]
  45. Tudu, B.; Nalajala, N.; Reddy, K.; Saikia, P.; Gopinath, C. Electronic integration and thin film aspects of Au–Pd/rGO/TiO2 for improved solar hydrogen generation. ACS Appl. Mater. Inter. 2019, 11, 32869–32878. [Google Scholar] [CrossRef]
  46. Agegnehu, A.; Pan, C.; Tsai, M.; Rick, J.; Su, W.; Lee, J.; Hwang, B. Visible light responsive noble metal-free nanocomposite of V-doped TiO2 nanorod with highly reduced graphene oxide for enhanced solar H2 production. Int. J. Hydrogen Energy 2016, 41, 6752–6762. [Google Scholar] [CrossRef]
  47. Wei, X.; Cao, J.; Fang, F. A novel multifunctional Ag and Sr2+ co-doped TiO2@rGO ternary nanocomposite with enhanced p-nitrophenol degradation, and bactericidal and hydrogen evolution activity. RSC Adv. 2018, 8, 31822–31829. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic representation of the formation of (a) rGO/TiO2 heterojunction assembly pho-tocatalysts; (b) X-ray diffraction (XRD) patterns, (c) UV–visible diffuse reflectance spectra, and (d) FTIR of TiO2 and rGO/TiO2, respectively.
Figure 1. Schematic representation of the formation of (a) rGO/TiO2 heterojunction assembly pho-tocatalysts; (b) X-ray diffraction (XRD) patterns, (c) UV–visible diffuse reflectance spectra, and (d) FTIR of TiO2 and rGO/TiO2, respectively.
Nanomaterials 12 01474 g001aNanomaterials 12 01474 g001b
Figure 2. The SEM images of (a,b) TiO2 and (c,d) rGO/TiO2, and the (e) TEM, (f,g) HRTEM images and (hj) elemental mappings of the rGO/TiO2 heterojunction assembly.
Figure 2. The SEM images of (a,b) TiO2 and (c,d) rGO/TiO2, and the (e) TEM, (f,g) HRTEM images and (hj) elemental mappings of the rGO/TiO2 heterojunction assembly.
Nanomaterials 12 01474 g002
Figure 3. (a) Photocatalytic H2 evolution rate of rGO, TiO2 and rGO/TiO2, (b) recycling tests of rGO/TiO2, (c) the Nyquist plots of electrochemical impedance under AM 1.5G of TiO2 and rGO/TiO2, and (d) the Mott–Schottky plots of TiO2.
Figure 3. (a) Photocatalytic H2 evolution rate of rGO, TiO2 and rGO/TiO2, (b) recycling tests of rGO/TiO2, (c) the Nyquist plots of electrochemical impedance under AM 1.5G of TiO2 and rGO/TiO2, and (d) the Mott–Schottky plots of TiO2.
Nanomaterials 12 01474 g003
Scheme 1. The mechanism diagram of photocatalytic hydrogen production of the rGO/TiO2 heterojunction assembly.
Scheme 1. The mechanism diagram of photocatalytic hydrogen production of the rGO/TiO2 heterojunction assembly.
Nanomaterials 12 01474 sch001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Z.; Yang, D.; Chu, H.; Guo, L.; Chen, T.; Mu, Y.; He, X.; Zhong, X.; Huang, B.; Zhang, S.; et al. Efficient Charge Transfer Channels in Reduced Graphene Oxide/Mesoporous TiO2 Nanotube Heterojunction Assemblies toward Optimized Photocatalytic Hydrogen Evolution. Nanomaterials 2022, 12, 1474. https://doi.org/10.3390/nano12091474

AMA Style

Li Z, Yang D, Chu H, Guo L, Chen T, Mu Y, He X, Zhong X, Huang B, Zhang S, et al. Efficient Charge Transfer Channels in Reduced Graphene Oxide/Mesoporous TiO2 Nanotube Heterojunction Assemblies toward Optimized Photocatalytic Hydrogen Evolution. Nanomaterials. 2022; 12(9):1474. https://doi.org/10.3390/nano12091474

Chicago/Turabian Style

Li, Zhenzi, Decai Yang, Hongqi Chu, Liping Guo, Tao Chen, Yifan Mu, Xiangyi He, Xueyan Zhong, Baoxia Huang, Shiyu Zhang, and et al. 2022. "Efficient Charge Transfer Channels in Reduced Graphene Oxide/Mesoporous TiO2 Nanotube Heterojunction Assemblies toward Optimized Photocatalytic Hydrogen Evolution" Nanomaterials 12, no. 9: 1474. https://doi.org/10.3390/nano12091474

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