Next Article in Journal
Benchmarking Acidic and Basic Catalysis for a Robust Production of Biofuel from Waste Cooking Oil
Next Article in Special Issue
Study on the Photocathodic Protection of Q235 Steel by CdIn2S4 Sensitized TiO2 Composite in Splash Zone
Previous Article in Journal
Removal of Banana Tree Fungi Using Green Tuff Rock Powder Waste Containing Zeolite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Facile and Large-scale Synthesis of Defective Black TiO2−x(B) Nanosheets for Efficient Visible-light-driven Photocatalytic Hydrogen Evolution

School of Materials Science and Technology, University of Shanghai for Science and Technology, Jungong Rd.516, Shanghai 200093, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to the work.
Catalysts 2019, 9(12), 1048; https://doi.org/10.3390/catal9121048
Submission received: 6 November 2019 / Revised: 6 December 2019 / Accepted: 9 December 2019 / Published: 10 December 2019
(This article belongs to the Special Issue TiO2 for Photocatalytic Applications)

Abstract

:
In the work, we firstly report the facile and large-scale synthesis of defective black TiO2−x(B) nanosheets via a dual-zone NaBH4 reduction method. The structure, physico-chemical, and optical properties of TiO2−x(B) nanosheets were systematically characterized by powder X-ray diffraction, Raman spectroscopy, UV-Vis absorption spectroscopy, and X-ray photoelectron spectroscopy, etc. The concentration of Ti3+ can be well tuned by NaBH4 reduction. With increasing the mass ratio of NaBH4 to TiO2(B), the generation of Ti3+ defects gives rise to the increased intensity of a broad band absorption in the visible wavelength range. It is demonstrated that the TiO2−x(B) photocatalyst synthesized with the mass ratio of NaBH4 to TiO2(B) of 3:1 exhibited an optimum photocatalytic activity and excellent photostability for hydrogen evolution under visible-light irradiation. By combining the advantages of 2D TiO2(B) nanosheets architecture with those of Ti3+ self-doping and simultaneous production of oxygen vacancy sites, the enhanced photocatalytic performance of the defective TiO2−x(B) nanosheets was achieved.

1. Introduction

Photocatalytic semiconducting materials for hydrogen evolution via water splitting have attracted considerable interest [1,2,3,4]. The anatase-, rutile-, and brookite-type TiO2 are the most widely studied photocatalysts. However, their potential applications are hindered severely by their large band gaps and consequently their limited visible-light-harvesting properties. Therefore, several approaches to enhance visible-light photoactivity and inhibit charge carrier recombination in TiO2-based photocatalysis have been developed, such as co-catalysts deposition [5,6], hetero/self-doping [7,8,9,10], junction composite [11,12], crystal facet engineering [13,14], surface disordering [15,16], etc. Especially, the intrinsic defects in the TiO2 matrix have been proved to trigger the visible-light activity of TiO2 [17,18,19,20,21,22]. The hydrogenation-induced defect-rich black TiO2−x displays remarkable stability and photoactivity. Meanwhile, the theoretical results also clearly demonstrate that a vacancy band state is formed as a result of the high vacancy concentration, thus achieving a narrow band gap (about 1.0 eV).
Apart from these TiO2 polymorphs, the crystalline structure of the metastable TiO2(B) nanosheet is found to be a layered and perovskite-like with lattice channels. The synthesis process is facile with ethylene glycol solutions and 140–180 °C, and only involves a one-step hydrolysis reaction of TiCl3 or TiCl4 [23,24]. Therefore, extensive efforts have been contributed to investigate the applications of the TiO2(B) nanosheet in the photocatalysis field [23,25,26,27,28]. However, due to the metastable structure and character of TiO2(B), almost no work has been reported on the facile and large-scale synthesis method and concurrently controllable structure and optical properties for defect-rich TiO2−x(B).
In this paper, we offer a facile dual-zone reduction approach to produce defect-rich TiO2−x(B) nanosheets by using NaBH4 as reductant agents. By adjusting the mass ratio of NaBH4/TiO2(B), the formation of Ti3+ and oxygen-vacancy defects can be well controlled. The as-synthesized defective TiO2−x(B) exhibited a broad absorption in the visible-light range, achieving the visible-light photoactivity for H2 evolution.

2. Results and Discussions

2.1. Fabrication of Defective Black TiO2−x(B) Nanosheets

The TiO2(B) phase commonly suffers from phase transformation into the stable TiO2 phase (anatase or rutile) in high-temperature annealing conditions, owing to the thermodynamically metastable structure. Thus, this work provides a new and simple dual-zone NaBH4 reduction approach to produce defective TiO2−x(B) nanosheets (Figure 1), showing the potential to replace the dangerous high-temperature hydrogenation method by directly using H2 as reductant reagent. Moreover, the byproducts residues from decomposition reactions of NaBH4 can be also avoided. Taking the advantage of the dual-zone reduction synthetic procedure, the phase transformation of TiO2(B) was inhibited, achieving self-doping of Ti3+ and simultaneous formation of oxygen vacancy sites.

2.2. Material Characterizations

Figure 2 shows the powder XRD patterns of the defective TiO2−x samples. The observed diffraction peaks at 2θ = 25.0°, 28.6°, and 48.6° can be assigned to the [110], [002], and [020] planes of the TiO2(B) phase (JCPDS No.74-1940), indicating that no phase transformation was observed on the as-synthesized TiO2−x(B) samples. The results show that under the mild reduction conditions, the simultaneous self-doping of Ti3+ and generation of oxygen vacancy sites were achieved, thus inhibiting the phase transformation of TiO2(B). In addition, the diffraction peaks intensity was decreased, not only indicating the decreased crystallinity of TiO2−x(B) along with increasing the mass ratio of NaBH4 to TiO2(B), but also shows that the TiO2(B) prepared by simple hydrothermal method possesses a low crystallinity and the reduction process at 200 °C further reduces the crystallinity.
Raman scatterings (shown as Figure 3) were measured to further examine the structure of the obtained TiO2−x(B) samples. No obvious change was observed on the Raman spectra of the as-synthesized TiO2−x(B) samples. However, with increasing the mass ratio of NaBH4 to TiO2(B), an obvious decrease in the Raman signal intensity was observed, indicating the lower crystallinity of the TiO2(B) phase, which is consistent with XRD results.
Figure 4 shows the UV-Vis absorption spectra (a) and Tauc plots (b) for the defective TiO2−x(B) nanosheets compared to the pristine TiO2(B). It can be seen that, with increasing the mass ratio of NaBH4 to TiO2(B), the UV-Vis absorption of defective TiO2−x(B) samples is enhanced to expand to the entire visible-light region, in accordance with distinct color change from white BT into black BT3 (inset of Figure 4b).The results confirm the assumption that a new vacancy band state, located below the conduction band minimum (CBM) of TiO2(B), can be formed by high concentration of Ti3+ doping. [17]
As shown in Figure 5a, the XPS survey spectra show the similar surface components of Ti and O elements in BT and BT3. The narrow scan XPS spectra of Ti 2p show that the binding energies of the spin doublet with Ti 2p3/2 and 2p1/2 are 458.4 and 464.1 eV, respectively (Figure 5b). The result indicates that the Ti species mainly exist as Ti4+, which is in good accordance with the literature results [29]. Moreover, the devolution of the Ti 2p XPS spectrum for the BT3 sample results in the obvious peaks belonging to Ti3+, for which the Ti 2p1/2 and Ti 2p3/2 peaks are located at about 457.5 and 463.2 eV, which confirms the generation of surface Ti3+ in TiO2(B). Such differences can be ascribed to the reduction of Ti4+ into Ti3+ after the NaBH4 treatment. As represented in Figure 5c, the deconvoluted peaks of O1s at 529.7, 531.0, and 532.8 eV are due to the lattice oxygen (Ti–O–Ti), surface hydroxyl group/O–defective matrix (Ti–O–H), and Ti–O–C groups, respectively [30,31,32]. The observation of Ti–O–C groups can be attributed to the surface-absorbed ethylene glycolate [26]. More interesting, the amount of lattice Ti–O–Ti bonds relative to surface Ti–O–H bonds/O–defective matrix was notably increased in BT3, as compared to those of BT. The results indicate the simultaneous production of oxygen vacancy sites [26,32].

2.3. Photocatalytic Activity for H2 Evolution

The amounts and rates of H2 evolution from aqueous methanol solution under visible-light were measured to represent the photocatalytic activity of the defect-containing TiO2−x(B) with photodeposition of 0.03 wt.% Rh as co-catalysts, as shown in Figure 6. No H2 gases were evolved from pristine BT. Interestingly, by an intermittent visible-light irradiation (every 30 min radiation followed by 30 min interval), the stable and continuous H2 evolution (nearly linear correlation between the evolved hydrogen amount with the irradiation time) was observed on the as-synthesized defective TiO2−x(B) samples (Figure 6a), indicating no significant deactivation of H2 evolution and excellent stability of defective TiO2−x(B) nanosheets. With increasing the mass ratio of NaBH4/BT, the H2 evolution from BT1 increased drastically, reached the maximum average rate of 0.58 μmol·g−1·h−1 for BT3, and then decreased (Figure 6b). The results confirm that the appropriate amount of self-doped Ti3+ defects along with the production oxygen vacancy sites are the key factors, leading to enhanced photocatalytic performance of the defective TiO2−x(B) nanosheets.

3. Experimental

3.1. Chemicals

Titanium tetrachloride (TiCl4, ≥99%) and Na3RhCl6 were purchased from Alfa Aesar. Ethylene glycol (EG, ≥99%), sodium borohydride (NaBH4, ≥99%), and methanol (MeOH, ≥99.9%) were purchased from Sigma-Aldrich.

3.2. Photocatalysts Preparation

A modified hydrolysis method was used to synthesize the TiO2(B) nanosheet powders [23]. Typically, a desired amount of deionized water was added into the pre-mixture of TiCl4 and ethylene glycol, and then heated at 150 °C for 6 h in a Teflon-lined stainless-steel autoclave. Finally, centrifugation, washing by deionized water and ethanol, and drying at 80 °C all night were performed to obtain products (denoted as BT).
The BT as precursor and the NaBH4 as reductant agent at different mass ratios of 1 to 4 were separately placed in a dual-zone quartz tube furnace. The reduction processes were carried out in an argon atmosphere with a flow rate of 10 mL/min (5.0 quality) at a heating rate of 5 °C/min. The NaBH4 and BT samples were annealed at 500 and 200 °C for 1 h, respectively. After being cooled to room temperature, the samples were thoroughly washed with deionized water and ethanol several times, and dried at 80 °C overnight. The obtained products were accordingly denoted as BT1, BT2, BT3, and BT4, with respect to the different mass ratio of 1 and 4.

3.3. Characterization

The X-ray diffraction (XRD) patterns of all samples were recorded using a PANalytical MPD diffractometer, with the radiation source of Cu-Kα (λ = 0.1541 nm) X-ray emission, and the scan range was set to 10° to 70° (2θ), with step of 0.05°. Raman spectra of all samples were collected using a Renishaw Raman microscope equipped with a 514 nm excitation laser. Taking BaSO4 as a reference, the UV-Vis absorption spectra were measured using a Shimadzu UV-2450 spectrophotometer. The X-ray photoelectron spectroscopy (XPS) measurements were carried out in Thermo ESCALAB 250XI System (ThermoFisher Scientific, Waltham, MA, USA), consisting of the Mg-Kα X-ray radiation source ( = 1253.6 eV) operating at 250 W (14 kV) and a high resolution hemisphere energy analyzer.
The base pressure of about 5 × 10−10 mbar was maintained in the measurement chamber. To obtain an overall energy resolution of 0.25 eV, the fixed transmission mode and pass energy of 93.9 eV were adopted during the measurements. The charging effects were compensated by a flood gun. A piece of carbon tape (Nisshin EM Co. Ltd, Tokyo, Japan) was used to manually mount samples in the sample holder. The XPS peak deconvolution was performed using the Casa software (Version 2.3.15 RUB license, Casa Software Ltd, Teignmouth, UK, 2009) with Shirley background subtraction and Gaussian–Lorentzian broadening function.

3.4. Photocatalytic Activity Tests

Photocatalytic hydrogen production reactions were conducted in a homemade eight-parallel multi-zone reaction system with air-tight quartz reactors. A 500 W mid-pressure Hg lamp was equipped as light irradiation source, with a water filter and a 420 nm cut-off filter. The evolved gases were determined by gas chromatography method using a thermal conductivity detector (GC7900, Techcomp Ltd., Beijing, China, MS-5A column and high-purity N2 (5.0 quality) as carrier gas). Typically, 30 mg powders were dispersed in 30 mL 10 vol.% aqueous methanol solution (MeOH), and then the ultrasonication was carried out for 10 min. Subsequently, for in-situ photodeposition of optimum amount of 0.03 wt.% Rh as co-catalyst, the Na3RhCl6 solution was added as precursor. Before irradiation, including the photocatalysts, the whole system was purged with N2 (5.0 quality) to remove air completely.

4. Conclusions

In summary, a facile and large-scale approach of a dual-zone NaBH4 reduction method was used for preparing defective black TiO2−x(B) nanosheets. We demonstrate that the mass ratio of NaBH4 to TiO2(B) plays a critical role in controlling the self-doped Ti3+ defects and simultaneously produced oxygen vacancy sites towards engineering the defective TiO2−x(B). The presence of Ti3+ and oxygen vacancy defects gives rise to the significantly increased intensity of the broad band absorption in the visible wavelength range. Under visible-light irradiation, the photocatalytic performance of defective TiO2−x(B) photocatalysts was greatly enhanced with excellent stability for hydrogen evolution, as compared to the non-photoactive pristine TiO2(B). The synthetic approach for synthesis of defective TiO2−x(B) shows great significance for developing a highly efficient catalytic system.

Author Contributions

Design of experiments and preparation and review of original manuscript were performed by J.X., J.Z., P.W., Z.C., H.H. and T.H.; supervision, final reviewing, and editing acquisition were performed by J.X., P.W. and X.W. All authors discussed the results and contributed to the final manuscript.

Acknowledgments

We greatly appreciate the financial supports from the National Natural Science Foundation of China (51572173, 51602197, 51771121, 51772297 and 51702212), Shanghai Municipal Science and Technology Commission (18511110600), Shanghai Academic/Technology Research Leader Program (19XD1422900), Shanghai Eastern Scholar Program (QD2016014), and USST Science and Technology Development Program (2019KJFZ014 and 2019KJFZ015).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yang, P.; Tarascon, J.M. Towards systems materials engineering. Nat. Mater. 2012, 11, 560–563. [Google Scholar] [CrossRef] [PubMed]
  2. Bard, A.J.; Fox, M.A. Artificial Photosynthesis: Solar Splitting of Water to Hydrogen and Oxygen. Acc. Chem. Res. 1995, 28, 141–145. [Google Scholar] [CrossRef]
  3. Wang, P.; Chen, P.; Kostka, A.; Marschall, R.; Wark, M. Control of Phase Coexistence in Calcium Tantalate Composite Photocatalysts for Highly Efficient Hydrogen Production. Chem. Mater. 2013, 25, 4739–4745. [Google Scholar] [CrossRef]
  4. Zhan, W.; Sun, L.; Han, X. Recent Progress on Engineering Highly Efficient Porous Semiconductor Photocatalysts Derived from Metal–Organic Frameworks. Nano-Micro Lett. 2019, 11, 1. [Google Scholar] [CrossRef] [Green Version]
  5. Ai, G.; Mo, R.; Li, H.; Zhong, J. Cobalt phosphate modified TiO2 nanowire arrays as co-catalysts for solar water splitting. Nanoscale 2015, 7, 6722–6728. [Google Scholar] [CrossRef]
  6. Yang, J.; Wang, D.; Han, H.; Li, C. Roles of Cocatalysts in Photocatalysis and Photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909. [Google Scholar] [CrossRef]
  7. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef]
  8. Zhang, Y.; Xing, Z.; Liu, X.; Li, Z.; Wu, X.; Jiang, J.; Li, M.; Zhu, Q.; Zhou, W. Ti3+ Self-Doped Blue TiO2(B) Single-Crystalline Nanorods for Efficient Solar-Driven Photocatalytic Performance. ACS Appl. Mater. Interfaces 2016, 8, 26851–26859. [Google Scholar] [CrossRef]
  9. Ran, P.; Jiang, L.; Li, X.; Zuo, P.; Li, B.; Li, X.; Cheng, X.; Zhang, J.; Lu, Y. Redox shuttle enhances nonthermal femtosecond two-photon self-doping of rGO–TiO2−x photocatalysts under visible light. J. Mater. Chem. 2018, 6, 16430–16438. [Google Scholar] [CrossRef]
  10. Choi, J.; Park, H.; Hoffmann, M.R. Effects of Single Metal-Ion Doping on the Visible-Light Photoreactivity of TiO2. J. Phys. Chem. C 2010, 114, 783–792. [Google Scholar] [CrossRef] [Green Version]
  11. Sun, B.; Zhou, W.; Li, H.; Ren, L.; Qiao, P.; Li, W.; Fu, H. Synthesis of Particulate Hierarchical Tandem Heterojunctions toward Optimized Photocatalytic Hydrogen Production. Adv. Mater. 2018, 30, 1804282. [Google Scholar] [CrossRef] [PubMed]
  12. Marschall, R. Semiconductor Composites: Strategies for Enhancing Charge Carrier Separation to Improve Photocatalytic Activity. Adv. Funct. Mater. 2014, 24, 2421–2440. [Google Scholar] [CrossRef]
  13. Liu, G.; Yu, J.C.; Lu, G.Q.; Cheng, H. Crystal facet engineering of semiconductor photocatalysts: Motivations, advances and unique properties. Chem. Commun. 2011, 47, 6763–6783. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, S.; Yu, J.; Jaroniec, M. Anatase TiO2 with Dominant High-Energy {001} Facets: Synthesis, Properties, and Applications. Chem. Mater. 2011, 23, 4085–4093. [Google Scholar] [CrossRef]
  15. Zhou, W.; Li, W.; Wang, J.; Qu, Y.; Yang, Y.; Xie, Y.; Zhang, K.; Wang, L.; Fu, H.; Zhao, D. Ordered Mesoporous Black TiO2 as Highly Efficient Hydrogen Evolution Photocatalyst. J. Am. Chem. Soc. 2014, 136, 9280–9283. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, X.; Liu, L.; Huang, F. Black titanium dioxide (TiO2) nanomaterials. Chem. Soc. Rev. 2015, 44, 1861–1885. [Google Scholar] [CrossRef] [PubMed]
  17. Fang, W.; Xing, M.; Zhang, J. A new approach to prepare Ti3+ self-doped TiO2 via NaBH4 reduction and hydrochloric acid treatment. Appl. Catal. B Environ. 2014, 160–161, 240–246. [Google Scholar] [CrossRef]
  18. Chen, X.; Lei, L.; Yu, P.Y.; Mao, S.S. Increasing Solar Absorption for Photocatalysis with Black Hydrogenated Titanium Dioxide Nanocrystals. Science 2011, 331, 746–750. [Google Scholar] [CrossRef]
  19. Ullattil, S.G.; Periyat, P. A ‘one pot’gel combustion strategy towards Ti3+ self-doped ‘black’anatase TiO2−x solar photocatalyst. J. Mater. Chem. A 2016, 4, 5854–5858. [Google Scholar] [CrossRef]
  20. Santara, B.; Giri, P.; Imakita, K.; Fujii, M. Evidence for Ti interstitial induced extended visible absorption and near infrared photoluminescence from undoped TiO2 nanoribbons: An in situ photoluminescence study. J. Phys. Chem. C 2013, 117, 23402–23411. [Google Scholar] [CrossRef]
  21. Naldoni, A.; Altomare, M.; Zoppellaro, G.; Liu, N.; Kment, S.; Zboril, R.; Schmuki, P. Photocatalysis with reduced TiO2: From black TiO2 to cocatalyst-free hydrogen production. ACS Catal. 2018, 9, 345–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Kumaravel, V.; Mathew, S.; Bartlett, J.; Pillai, S.C. Photocatalytic hydrogen production using metal doped TiO2: A review of recent advances. Appl. Catal. B Environ. 2018, 244, 1021–1064. [Google Scholar] [CrossRef]
  23. Xiang, G.; Li, T.; Zhuang, J.; Wang, X. Large-scale synthesis of metastable TiO2(B) nanosheets with atomic thickness and their photocatalytic properties. Chem. Commun. 2010, 46, 6801–6803. [Google Scholar] [CrossRef] [PubMed]
  24. Zhenyu, S.; Xing, H.; Martin, M.; Wolfgang, S.; Edgar, V. A carbon-coated TiO2(B) nanosheet composite for lithium ion batteries. Chem. Commun. 2014, 50, 5506–5509. [Google Scholar]
  25. Liu, P.; Zhao, Y.; Qin, R.; Mo, S.; Chen, G.; Gu, L.; Chevrier, D.M.; Zhang, P.; Guo, Q.; Zang, D.; et al. Photochemical route for synthesizing atomically dispersed palladium catalysts. Science 2016, 352, 797–800. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, P.; Yi, Z.; Zhang, J.; Cai, Z.; Lyu, B.; Yang, J.; Wang, X. In-situ photosynthetic route to tailor point defects in TiO2(B) nanosheets for visible light-driven photocatalytic hydrogen production. ChemCatChem 2019, 11, 4252–4255. [Google Scholar] [CrossRef]
  27. Kong, X.; Xu, Y.; Cui, Z.; Li, Z.; Liang, Y.; Gao, Z.; Zhu, S.; Yang, X. Defect enhances photocatalytic activity of ultrathin TiO2(B) nanosheets for hydrogen production by plasma engraving method. Appl. Catal. B Environ. 2018, 230, 11–17. [Google Scholar] [CrossRef]
  28. Wan, N.; Xing, Z.; Kuang, J.; Li, Z.; Yin, J.; Zhu, Q.; Zhou, W. Oxygen vacancy-mediated efficient electron-hole separation for CNS-tridoped single crystal black TiO2(B) nanorods as visible-light-driven photocatalysts. Appl. Surf. Sci. 2018, 457, 287–294. [Google Scholar] [CrossRef]
  29. Gai, Z.; Cheng, Z.; Wang, X.; Zhao, L.; Yin, N.; Abah, R.; Zhao, M.; Hong, F.; Yu, Z.; Dou, S. A colossal dielectric constant of an amorphous TiO2:(Nb, In) film with low loss fabrication at room temperature. J. Mater. Chem. C 2014, 2, 6790–6795. [Google Scholar] [CrossRef]
  30. Zhang, P.; Shao, C.; Zhang, Z.; Zhang, M.; Mu, J.; Guo, Z.; Liu, Y. TiO2@carbon core/shell nanofibers: Controllable preparation and enhanced visible photocatalytic properties. Nanoscale 2011, 3, 2943–2949. [Google Scholar] [CrossRef]
  31. Huang, N.-P.; Michel, R.; Voros, J.; Textor, M.; Hofer, R.; Rossi, A.; Elbert, D.L.; Hubbell, J.A.; Spencer, N.D. Poly (L-lysine)-g-poly (ethylene glycol) layers on metal oxide surfaces: Surface-analytical characterization and resistance to serum and fibrinogen adsorption. Langmuir 2001, 17, 489–498. [Google Scholar] [CrossRef]
  32. Das, J.; Pradhan, S.; Sahu, D.; Mishra, D.; Sarangi, S.; Nayak, B.; Verma, S.; Roul, B. Micro-Raman and XPS studies of pure ZnO ceramics. Phys. B Condens. Matter 2010, 405, 2492–2497. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration for the formation of defective black TiO2−x(B) nanosheets.
Figure 1. Schematic illustration for the formation of defective black TiO2−x(B) nanosheets.
Catalysts 09 01048 g001
Figure 2. Powder XRD patterns of the as-synthesized pristine TiO2(B) and defective TiO2−x(B) nanosheets.
Figure 2. Powder XRD patterns of the as-synthesized pristine TiO2(B) and defective TiO2−x(B) nanosheets.
Catalysts 09 01048 g002
Figure 3. Raman spectra of defective TiO2−x(B) nanosheets in comparison to that of pristine TiO2(B).
Figure 3. Raman spectra of defective TiO2−x(B) nanosheets in comparison to that of pristine TiO2(B).
Catalysts 09 01048 g003
Figure 4. UV-Vis absorption spectra (a) and Tauc plots (b) of pristine TiO2(B) and defective TiO2−x(B) nanosheets. Inset of (b), distinct color change from white BT to black BT3.
Figure 4. UV-Vis absorption spectra (a) and Tauc plots (b) of pristine TiO2(B) and defective TiO2−x(B) nanosheets. Inset of (b), distinct color change from white BT to black BT3.
Catalysts 09 01048 g004
Figure 5. XPS spectra of pristine TiO2(B) and defective TiO2−x(B): Survey scans (a), narrow scan Ti 2p (b), narrow scan O 1s (c).
Figure 5. XPS spectra of pristine TiO2(B) and defective TiO2−x(B): Survey scans (a), narrow scan Ti 2p (b), narrow scan O 1s (c).
Catalysts 09 01048 g005
Figure 6. (a) Photocatalytic H2 evolution over the as-synthesized defective TiO2−x(B) nanosheets under visible-light irradiation (>420 nm) and reaction conditions—0.03 g of catalysts, 30 mL of 10 vol.% aqueous methanol solution, loading of 0.03 wt.% Rh as co-catalysts. (b) The corresponding average H2 evolution rates of the as-synthesized defective TiO2−x(B). Error bars are generated by measurements repeated at least three times for all samples, with less than 5% deviations for the samples prepared in different batches.
Figure 6. (a) Photocatalytic H2 evolution over the as-synthesized defective TiO2−x(B) nanosheets under visible-light irradiation (>420 nm) and reaction conditions—0.03 g of catalysts, 30 mL of 10 vol.% aqueous methanol solution, loading of 0.03 wt.% Rh as co-catalysts. (b) The corresponding average H2 evolution rates of the as-synthesized defective TiO2−x(B). Error bars are generated by measurements repeated at least three times for all samples, with less than 5% deviations for the samples prepared in different batches.
Catalysts 09 01048 g006

Share and Cite

MDPI and ACS Style

Xu, J.; Zhang, J.; Cai, Z.; Huang, H.; Huang, T.; Wang, P.; Wang, X. Facile and Large-scale Synthesis of Defective Black TiO2−x(B) Nanosheets for Efficient Visible-light-driven Photocatalytic Hydrogen Evolution. Catalysts 2019, 9, 1048. https://doi.org/10.3390/catal9121048

AMA Style

Xu J, Zhang J, Cai Z, Huang H, Huang T, Wang P, Wang X. Facile and Large-scale Synthesis of Defective Black TiO2−x(B) Nanosheets for Efficient Visible-light-driven Photocatalytic Hydrogen Evolution. Catalysts. 2019; 9(12):1048. https://doi.org/10.3390/catal9121048

Chicago/Turabian Style

Xu, JingCheng, JiaJia Zhang, ZhengYang Cai, He Huang, TianHao Huang, Ping Wang, and XianYing Wang. 2019. "Facile and Large-scale Synthesis of Defective Black TiO2−x(B) Nanosheets for Efficient Visible-light-driven Photocatalytic Hydrogen Evolution" Catalysts 9, no. 12: 1048. https://doi.org/10.3390/catal9121048

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