Next Article in Journal
Effect of Dopant Loading on the Structural and Catalytic Properties of Mn-Doped SrTiO3 Catalysts for Catalytic Soot Combustion
Previous Article in Journal
New Trendy Magnetic C-Scorpionate Iron Catalyst and Its Performance towards Cyclohexane Oxidation
Previous Article in Special Issue
Synthesis of Ag3PO4/G-C3N4 Composite with Enhanced Photocatalytic Performance for the Photodegradation of Diclofenac under Visible Light Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Low Temperature Synthesis of Nest-Like Microsphere with Exposed (001) Facets and Its Enhanced Photocatalytic Performance by NaOH Alkalization

College of Geology and Environment, Xi’an University of Science and Technology, Xi’an 710054, China
*
Author to whom correspondence should be addressed.
Catalysts 2018, 8(2), 70; https://doi.org/10.3390/catal8020070
Submission received: 31 December 2017 / Revised: 5 February 2018 / Accepted: 5 February 2018 / Published: 8 February 2018
(This article belongs to the Special Issue Organic Photoredox Catalysis)

Abstract

:
In this study, we completed a simple low-temperature synthesis of nest-like titanium oxide (TiO2) microspheres with exposed (001) facets. For the first time, the photocatalytic performance was enhanced by sodium hydroxide (NaOH) alkalization. The characterization of as-synthesized F-TiO2 and OH-TiO2 were analyzed by field emission scanning electron microscopy, high-resolution transmission electron microscopy, X-ray diffraction, fourier transform infrared spectroscopic analysis, ultraviolet-vis diffuse reflection spectra and Raman spectroscopy. The photocatalytic activity of the as-prepared catalyst was evaluated through the photocatalytic degradation of methylene blue (MB) and Rhodamine B (RhB) under simulated solar light. The results showed that modification using NaOH can lead to an increase in the percentage of (001) facets from 27.8% for F-TiO2 to 39.2% for OH-TiO2. OH-TiO2 showed superior catalytic photoactivity toward MB. The mechanism of NaOH on TiO2 is also discussed.

Graphical Abstract

1. Introduction

Since Fujishima discovered that titanium oxide (TiO2) could split water in 1972 [1], TiO2 has been studied for its lack of toxicity, low cost and stability, which make it a promising photocatalyst for environmental remediation [2,3,4] and hydrogen evolving [5]. The decreasing order of the surface energies of anatase TiO2 has been found to be 0.90 J/m2 for (001), 0.53 J/m2 for (100) and 0.44 J/m2 for (101) [6,7], so most anatase TiO2 has a low-energy (101) surface rather than high-energy (001) facets [8,9,10,11,12,13,14,15,16]. The (001) surface of anatase TiO2 is much more reactive than the (101) surface [17,18]. Developing a technique to exposure the surface of (001) has proven to be a challenge.
In 2008, Yang et al. successfully prepared anatase TiO2 crystals with as much as 47% of the (001) facets being exposed using a surface fluorination method [19]. Since then, studies on (001) faceted anatase TiO2 have been completed [20,21,22,23,24,25,26,27]. However, most of the (001) faceted anatase TiO2 crystals were prepared at high temperature through calcination or hydrothermal techniques [20,21,22,23,24,25,26,27].
Akali-modification of catalysts have also been proven to be effective on enhancing the photo-activity in methane dehydroaromatization, cumene cracking and carbon monoxide (CO) oxidation [28,29,30]. For example, Han et al. [28] found that alkali could form more hydroxyl on gold (Au) catalysts to enhance their catalytic activity. Alkali modified ZSM-5 zeolite also showed enhanced catalytic performance due to the formation of additional mesopores and the improvement of mass transfer and reaction kinetics [29,30]. The performance of sodium hydroxide (NaOH)-modified Pt/TiO2 was enhanced with the oxidation of formaldehyde at room temperature [30]. This present study attempted to modify TiO2/TiOF2 by NaOH to enhance its catalytic performance.
Furthermore, F-doped TiO2 is thought to have (001) facet exposure, and the fluoride ions remain well bound and stable on the surface of F-TiO2 samples. Further study is required on the mechanism by which the NaOH modification occurs on F-TiO2, the morphology, (001) facet exposure, crystal properties, ultraviolet (UV)-vis adsorption properties, and the crystal defect changes of F-TiO2 after NaOH-modification. In this paper, a low-temperature synthesis method for TiO2 microspheres with exposed (001) facets is reported, and the photocatalytic performance was enhanced by NaOH alkalization (nest-like OH-TiO2) for the first time. The morphology, (001) facet exposure, crystal properties, UV-vis adsorption properties, and crystal defects changes were studied. The photocatalytic performance of the as-prepared catalyst was tested using the photocatalytic degradation of methylene blue (MB) and Rhodamine B (RhB) under simulated solar light. The NaOH alkalized microspheres showed superior catalytic photoactivity towards MB. The mechanism of NaOH toward TiO2 is also discussed.

2. Results and Discussion

2.1. Structural Characterization

Figure 1 shows the X-ray diffraction (XRD) patterns of as-prepared F-TiO2 and OH-TiO2 samples. No apparent peaks of other phases were found, meaning that both samples were pure crystals. The XRD peak positions for both the samples at 2θ were 25.28°, 37.80°, 48.04°, 53.89°, 55.06°, 62.68°, 70.31° and 75.03°, which matches well with anatase TiO2 (JCPDS card No. 21-1272) and different from P25 (mixed crystals of anatase and rutile TiO2). The characteristic diffraction peak intensity of F-TiO2 is obviously stronger than that of OH-TiO2, meaning that the crystallinity of the sample can be increased with fluorine ions, which aligns with the results of Xiang et al. [31]. After NaOH treatment, the crystal form of the catalyst changed, which aligns with the results of Hou Chentao et al. [32].
The morphology of the as-prepared samples was characterized by field emission scanning electron microscopy (FE-SEM), and high-resolution transmission electron microscopy (HRTEM) in Figure 2. The as-synthesized F-TiO2 has a hollow microsphere shape with an average size of around 700–900 nm covered with compact particles of around 50 nm (Figure 2a). This shape may be explained by the Ostwald ripening process [33]. In the HRTEM (Figure 2c and, especially in Figure 2d), the spacing of the lattice fringe was 0.352 nm and 0.235 nm, respectively, corresponding to the lattice distance of the (101) and (001) plane of anatase TiO2 [34,35]. So the surfaces of the F-TiO2 microspheres are covered with small (001) exposure crystals. From Figure 2b, after modification by NaOH, the microspheres were destroyed and the compact particles became loose and hairy, developing a nest-like structure. As can be seen in Figure 2e,f, the nest-like structure also has the (001) surface exposed, permitting light-scattering inside the shell wall and enhancing absorption in light. We think the mechanics may be as follows. When the NaOH is added, the F ions will be exchanged by OH of NaOH, the transformation takes place starting from the edges and corners of the F-TiO2 [36], which are of high surface energy. The transformation leads to the gradual formation of OH-TiO2 with more (001) facets exposure and flatter TiO2 sheets. The transformation also makes the compact F-TiO2 crystals looser and take on a nest-like shape. The nest-like structure increases the OH-TiO2 surface area (SBET) to 55 m2/g, which is much larger than that of F-TiO2 at 20 m2/g.
Raman spectroscopy was used to calculate the percentage of (001) facets of the as-prepared catalysts (Figure 3). The percentage of (001) facets was calculated using the peak intensity ratio of the Eg and A1g peaks at 144 cm1 and 514 cm1, respectively [27,37]. The percentage of (001) facets for F-TiO2 was 27.8%, but after modification with NaOH, the percentage of OH-TiO2 increased to 39.2%, which helped to enhance the photocatalytic performance of catalyst.

2.2. UV-Vis DRS Analysis

The UV-vis diffuse reflectance spectra (DRS) and bandgap of the as-prepared samples are shown in Figure 4. From Figure 4a, the F-TiO2 and OH-TiO2 catalysts of are almost identical in terms of absorption between 230 and 550 nm. In the range of 550 to 850 nm, OH-TiO2 has a light stronger absorption than F-TiO2, showing that OH-TiO2 exhibits stronger light absorption ability in the visible light range. From Figure 4b, the OH-TiO2 bandgap is 3.09 eV, which is somewhat smaller than that of F-TiO2 (3.17 eV). This may be attributed to the nest-like microstructure morphology of OH-TiO2, which may promote light-scattering inside the pores of the shell wall and enhance absorption in light [25].

2.3. FT-IR Analysis

The Fourier transform infrared (FT-IR) spectroscopy was employed to investigate the chemical bonding of the catalysts (Figure 5a). The absorption peaks of F-TiO2 and OH-TiO2 were observed at 3416 cm−1 and 1630 cm−1, respectively, which could be assigned to the adsorbed and bound H2O on the particles, respectively [38,39,40]. The span between 2934 and 3416 cm−1 is attributed to the O–H bonds or associated O–H. O–H bonds appeared both in F-TiO2 and OH-TiO2 [39,40]. The peaks around 672 cm−1 correspond to the characteristic absorption of Ti–O [29]. The peaks around 930 cm−1 are attributed to F–Ti which disappeared after NaOH washing, meaning F was exchanged by OH.

2.4. Photoluminescence Analysis

The Photoluminescence (PL) emission spectra were used to investigate the efficiency of charge carrier trapping, immigration and transfer, and to understand the fate of electron hole pairs in catalysts (Figure 5b). Six peaks were observed in the spectra. The broad emission band centered at 396 nm (peak 1) and 407 nm (peak 2) is ascribed to bound exciton emission due to the trapping of free excitons by titanate groups near defects [41]. The long wavelength range, from 440 to 465 nm (peaks 3, 4 and 5), is attributed to the oxygen vacancy with two trapped electrons. Peak 6, at 525 nm, is assigned to the oxygen vacancy as a result of the Franck–Condon principle and the polarizability of the lattice ions surrounding the vacancy or one trapped electron, i.e., with a Ti3+ or F+ center [41]. OH-TiO2 has more oxygen vacancies than F-TiO2. Oxygen vacancy sites are important for the formation of superoxide (O2·) and hydroxyl (·OH) radicals for photocatalytic degradation.

2.5. Photocatalytic Activity

In the present work, the photocatalytic activity of F-TiO2 and OH-TiO2 was evaluated by monitoring the degradation of methylene blue (MB) and Rhodamine B (RhB) solution under simulated solar irradiation. The degradation performance of catalysts and the variation in the UV-vis absorption spectra of the MB solution exposed to different TiO2 samples are shown in Figure 6. From Figure 6a, after modification with NaOH, the degradation performance of TiO2 was considerably enhanced, especially for MB. The performance was almost the same as or better than TiO2 samples synthesized under high temperature or samples having other morphology [14,42]. In Figure 6b, from the line of no light, 20 mg, C/C0 almost no longer decreased after 0.5 h, meaning that it is an adsorption process in this 0.5 h, and the adsorption balance is reached after 0.5 h. From the line representing no catalysts, we can see that light played little role on the discoloration. Therefore, we can conclude that with catalysts the discoloration process is adsorption process in dark for 0.5 h, after that the discoloration process is a photocatalytic degradation.
Figure 6c,d shows the UV-vis absorption spectral changes of the MB solution by OH-TiO2 and F-TiO2, respectively. According to previous research, two processes can cause the discoloration of MB: oxidative degradation and two-electron reduction to leuco-MB which can be detected by the UV-vis absorption at 256 nm [43,44]. From Figure 6d, a blue-shift from 665 to 625 nm and an absorbent peak at 256 nm emerged in the spectral change of MB with irradiation on OH-S0.5. This means that reductive conversion to leuco-MB occurs in the degradation path of MB.

2.6. Radical-Scavenging Experiments

Radical-scavenging experiments were performed to complete an in-depth study of the photocatalytic degradation mechanism (Figure 7). Terephthalic acid can combine with the hydroxyl radicals (·OH) to reduce the activity of the catalyst, and the benzoquinone can combine with the superoxide radical (O2·) to decrease the activity of the catalyst. When terephthalic acid was added to the degradation system, the efficiency decreased more than with benzoquinone when benzoquinone was added, demonstrating that the hydroxyl radicals play a leading role in photocatalysis.

3. Materials and Methods

3.1. Materials

Tetrabutyl titanate, hydrofluoric acid, methylene blue, and benzoquinone (AR, KeLong Co., Ltd., Chengdu, China), anhydrous ethanol (AR, ChuanDong Co., Ltd., Chongqing, China), sodium hydroxide (AR, Kermel Co., Ltd., Tianjin, China), Rhodamine B (AR, FuChen Co., Ltd., Tianjin, China), barium sulfate (AR, XiLong Co., Ltd., Chengdu, China), and terephthalic acid (AR, Shanghai Macklin Biochemical Co., Ltd., Shanghai, China) were used for the experiments.

3.2. Preparation of Catalysts

A total of 15.20 mL anhydrous ethanol was added to 17.60 mL tetrabutyl titanate, denoted as solution A; then15.20 mL anhydrous ethanol was added to 90 mL ultra-pure water, with 6 mL HF, recorded as solution B; solution A was added dropwise into solution B with a dropping speed of 2–3 drops per second and mixed at a low speed for 2 h, the obtained titanium dioxide gel was stored at room temperature for two days. The aged TiO2 gel was moved to a stainless-steel reactor containing polytetrafluoroethylene, and kept at constant temperature of 100 °C for 2 h. After cooling, the as-prepared power was washed with deionized water and ethanol 3 times and dried in a 100 °C drying oven. After grinding, white powder was obtained, which was F-TiO2.
Half of the above F-TiO2 powder was mixed with 10 g NaOH and added to 100 mL distilled water, and stirred at medium speed for 2 h, then centrifugally settled, washed 3 times with deionized water, dried in a 100 °C drying oven, and grinded to obtain the OH-TiO2 samples.

3.3. Catalyst Characterization

The crystal structure of the powders were analyzed by X-ray diffractometry (XRD) (Bruker, D8-Advance, Rheinstetten, Germany) with Cu Kα radiation (λ = 0.154 nm, 40 kV, 40 mA) with a scanning range of 20° to 80°. The UV-Vis absorption spectra of the powders were obtained from dry-pressed disk samples with a UV-Vis spectrometer (TU-1911, Shanghai, China), BaSO4 was used as a reflectance standard in the UV-Vis diffuse reflectance. The microstructure of the powers were examined by field-emission scanning (FE-SEM) (JEOL, JSM-6700F, Tokyo, Japan) operating at 5.0 KV and transmission electron microscopy and high resolution transmission electron microscopy (HRTEM) (Tecnai G2 F20, FEI, Hillsboro, OR, USA), using a 200 kV accelerating voltage. Fourier transform infrared (FT-IR) spectra were analyzed (Bruker, Tensor27, Rheinstetten, Germany) in the wavenumber range from 4000 to 400 cm−1. The Photoluminescence (PL) emission spectra were measured at room temperature with a fluorescence spectrophotometer (Hitachi F-2700, Tokyo, Japan) using 325 nm line with a Xe lamp.

3.4. Photocatalytic Activity Tests

The photocatalytic activities of both samples were evaluated by the degradation of methylene blue (MB) and Rhodamine B (RhB) solution under simulated solar light. The photocatalytic degradation tests were performed in an 200 mL double-layered quartz glass reactor With either 100 mL 10 mg/L MB or RhB and certain quantities of catalysts. Cooling water was introduced into the interlayer of the quartz reactor to maintain the solution at room temperature. A Jiguang-300W Xe lamp (simulating solar light) was located 30 cm away from the MB or RhB solution. When the solution was magnetically stirred for 0.5 h in dark to obtain the adsorption-desorption equilibrium, the Xe lamp was turned on to initiate the degradation. About 4.0 mL of solution was extracted and centrifuged at a speed of 11,000 r/min to remove catalysts every 0.5 h. Then, the MB or RhB concentration were analyzed on a Purkinje UV1901 UV-vis spectrophotometer at 665 nm or 554 nm. The photocatalyst was separated from the MB or RhB solution and another run was started to investigate the catalysts durability. Because the absorbance of simulated azo dye wastewater has a linear relationship with the concentration of dye contained therein, the degradation of dyes is indirectly reflected by measuring the absorbance change of simulated dye wastewater.
The photocatalytic reaction efficiency (100%) = A/A0 × 100% = C/C0 × 100%.
Where A0 is the absorbance before the reaction, and A is the absorbance of every 0.5 h centrifuge in the reaction; C0 is the absorbance before the reaction, and C is the absorbance obtained by every 0.5 h centrifuge in the reaction.
Terephthalic acid (3 mmol/L) and p-benzoquinone were added to a mixed solution containing 20 mg of OH-TiO2 and 100 mL of 10 mg/L of methylene blue solution, and methylene blue was degraded as a control.

4. Conclusions

A simple low-temperature synthesis of nest-like TiO2 microspheres with exposed (001) facets was reported and its photocatalytic performance was enhanced by NaOH alkalization for the first time in this study. The photocatalytic activity of the as-prepared catalyst was evaluated by photocatalytic degradation of methylene blue and Rhodamine B under simulated solar light. The results show that the NaOH alkalized microspheres had superior catalytic photoactivity compared to (001)TiO2 towards MB. The reason for this result may be related to its unique nest-like morphology, increased (001) facet exposure, more O–H, and more oxygen vacancy sites. Radical-scavenging experiments revealed that the hydroxyl radicals play a leading role in photocatalysis.

Acknowledgments

Financial support was provided by Shaanxi key industrial projects (2014GY2-07) and Shaanxi Province education department science and technology research plan (15JK1460).

Author Contributions

In this paper, Chentao Hou and Jiaming Zhu designed the experiments; Jiaming Zhu and Qiaoqiao Song conducted the experiments; Chentao Hou and Jiaming Zhu analyzed the data; Chentao Hou wrote the article.

Conflicts of Interest

No conflict of interest exists in this manuscript submission, and it has been approved by all of the authors for publication. All authors listed have approved the manuscript enclosed.

References

  1. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  2. Tian, F.; Wu, Z.; Chen, Q.; Yan, Y.; Cravotto, G.; Wu, Z. Microwave-induced crystallization of AC/TiO2 for improving the perfomnance of rhodamine B dye degradation. Appl. Surf. Sci. 2015, 351, 104–112. [Google Scholar] [CrossRef]
  3. Atout, H.; Álvarez, M.G.; Chebli, D.; Bouguettoucha, A.; Tichit, D.; Llorca, J.; Medina, F. Enhanced photocatalytic degradation of methylene blue: Preparation of TiO2/reduced graphene oxide nanocomposites by direct sol-gel and hydrothermal methods. Mater. Res. Bull. 2017, 50, 952–957. [Google Scholar] [CrossRef]
  4. Zhang, S.T.; Ruan, Y.R.; Liu, C.; Wang, P.; Ma, Y.Q. The evolution of structure, chemical state and photocatalytic performance of α-Fe/FeTiO3/TiO2 with the nitridation at different temperatures. Mater. Res. Bull. 2017, 95, 503–508. [Google Scholar] [CrossRef]
  5. Bessegato, G.G.; Guaraldo, T.T.; Brito, J.F.; Brugnera, M.F. Achievements and trends in photoelectrocatalysis: From environmental to energy applications. Electrocatalysis 2015, 6, 415–441. [Google Scholar] [CrossRef]
  6. Michele, L.; Andrea, V.; Annabella, S. Structure and energetics of stoichiometric TiO2 anatase surfaces. Phys. Rev. B 2001, 65, 155409. [Google Scholar] [CrossRef]
  7. Minero, C.; Mariella, G.; Maurino, V.; Vione, D.; Pelizzetti, E. Photocatalytic transformation of organic compounds in the presence of inorganic ions Competitive reactions of phenol and alcohols on a titanium dioxide-fluoride system. Langmuir 2000, 16, 8964–8972. [Google Scholar] [CrossRef]
  8. Huang, W.C.; Ting, J.-M. Novel nitrogen-doped anatase TiO2 mesoporous bead photocatalysts for enhanced visible light response. Ceram. Int. 2017, 43, 9992–9997. [Google Scholar] [CrossRef]
  9. Khalid, N.R.; Majid, A.; Tahir, M.B.; Niaz, N.A.; Khalid, S. Carbonaceous-TiO2 nanomaterials for photocatalytic degradation of pollutants: A review. Ceram. Int. 2017, 43, 14552–14571. [Google Scholar] [CrossRef]
  10. Gjipalaj, J.; Alessandri, I. Easy recovery; mechanical stability; enhanced adsorption capacity and recyclability of alginate-based TiO2 macrobead photocatalysts for water treatment. J. Environ. Chem. Eng. 2017, 5, 1763–1770. [Google Scholar] [CrossRef]
  11. Chowdhury, I.H.; Ghosh, S.; Basak, S.; Naskar, M.K. Mesoporous CuO-TiO2 microspheres for efficient catalytic oxidation of CO and photodegradation of methylene blue. J. Phys. Chem. Solids 2017, 104, 103–110. [Google Scholar] [CrossRef]
  12. Cao, X.; Luo, S.; Liu, C.; Chen, J. Synthesis of Bentonite-Supported Fe2O3-Doped TiO2 superstructures for highly promoted photocatalytic activity and recyclability. Adv. Powder Technol. 2017, 28, 993–999. [Google Scholar] [CrossRef]
  13. Liu, S.; Yu, J.; Jaroniec, M. Tunable Photocatalytic Selectivity of Hollow TiO2 Microspheres Composed of Anatase Polyhedra with Exposed {001} Facets. J. Am. Chem. Soc. 2010, 132, 11914–11916. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, J.; Li, W.; Li, X.; Zhang, X. Low temperature synthesis of water dispersible F-doped TiO2 nanorods with enhanced photocatalytic activity. RSC Adv. 2017, 7, 21547–21555. [Google Scholar] [CrossRef]
  15. Zhu, Z.F.; Zhou, J.Q.; He, Z.L.; Li, J.Q.; Liu, H. Preparation, characterisation and activity of TiO2−xFx spherical photocatalyst: Influence of sodium fluoride on methyl orange degradation. Mater. Res. Innov. 2011, 15, 78–82. [Google Scholar] [CrossRef]
  16. Wang, Q.; Chen, C.; Zhao, D.; Ma, W.; Zhao, J. Change of Adsorption Modes of Dyes on Fluorinated TiO2 and Its Effect on Photocatalytic Degradation of Dyes under Visible Irradiation. Langmuir 2008, 24, 7338–7345. [Google Scholar] [CrossRef] [PubMed]
  17. Yu, J.C.; Ho, W.K.; Yu, J.G.; Hark, S.K.; Iu, K. Effects of trifluoroacetic acid modification on the surface microstructures and photocatalytic activity of mesoporous TiO2 thin films. Langmuir 2003, 19, 3889–3896. [Google Scholar] [CrossRef]
  18. Lewandowski, M.; Ollis, D.F. Halide acid pretreatments of photocatalysts for oxidation of aromatic air contaminants: rate enhancement, rate inhibition and a thermodynamic rationale. Catalysis 2003, 217, 38–46. [Google Scholar] [CrossRef]
  19. Yang, H.G.; Sun, C.H.; Qiao, S.Z.; Zou, J.; Liu, G.; Smith, S.C.; Cheng, H.M.; Lu, G.Q. Anatase TiO2 single crystals with a large percentage of reactive facets. Nature 2008, 453, 638–641. [Google Scholar] [CrossRef] [PubMed]
  20. Peng, J.-D.; Lin, H.-H.; Lee, C.-T.; Tseng, C.-M.; Suryanarayanan, V.; Vittala, R.; Ho, K.-C. Hierarchically assembled microspheres consisting of nanosheets of highly exposed (001)-facets TiO2 for dye-sensitized solar cells. RSC Adv. 2016, 6, 14178–14191. [Google Scholar] [CrossRef]
  21. Tian, F.; Zhang, Y.; Zhang, J.; Pan, C. Raman Spectroscopy: A New Approach to Measure the Percentage of Anatase TiO2 Exposed (001) Facets. J. Phys. Chem. C 2012, 116, 7515–7519. [Google Scholar] [CrossRef]
  22. He, Z.; Wen, L.; Wang, D.; Xue, Y.; Lu, Q.; Wu, C.; Chen, J.; Song, S. Photocatalytic Reduction of CO2 in Aqueous Solution on Surface-Fluorinated Anatase TiO2 Nanosheets with Exposed {001} Facets. Energy Fuels 2014, 28, 3982–3993. [Google Scholar] [CrossRef]
  23. He, Z.; Cai, Q.; Hong, F.; Jiang, Z.; Chen, J.; Song, S. Effective Enhancement of the Degradation of Oxalic Acid by Catalytic Ozonation with TiO2 by Exposure of {001} Facets and Surface Fluorination. Ind. Eng. Chem. Res. 2012, 51, 5662–5668. [Google Scholar] [CrossRef]
  24. Nguyen, V.-H.; Lasek, J.; Jeffrey, C.S.; Yu, J.C.-C. Titania nanosheet photocatalysts with dominantly exposed (001) reactive facets for photocatalytic NOx abatement. Appl. Catal. B Environ. 2017, 219, 391–400. [Google Scholar] [CrossRef]
  25. Zhang, G.; Zhang, S.; Wang, L.; Liu, R.; Zeng, Y.; Xia, X.; Liu, Y.; Luo, S. Facile synthesis of bird’s nest-like TiO2 microstructure with exposed (001) facets for photocatalytic degradation of methylene blue. Appl. Surf. Sci. 2017, 391, 228–235. [Google Scholar] [CrossRef]
  26. Zhang, Y.; Xia, T.; Shang, M.; Wallenmeyer, P.; Katelyn, D.; Peterson, A.; Murowchick, J.; Dong, L.; Chen, X. Structural evolution from TiO2 nanoparticles to nanosheets and their photocatalytic performance in hydrogen generation and environmental pollution removal. RSC Adv. 2014, 4, 16146–16152. [Google Scholar] [CrossRef]
  27. Chu, L.; Qin, Z.; Yang, J.; Li, X. Anatase TiO2 Nanoparticles with Exposed {001} Facets for Efficient Dye-Sensitized Solar Cells. Sci. Rep. 2015, 5, 12143. [Google Scholar] [CrossRef] [PubMed]
  28. Han, D.; Zhou, C.; Yin, H.; Zhang, D.; Xu, X. Reactivity of the alkaline pretreated nanoporous gold for the CO oxidation. Catal. Lett. 2011, 141, 1026–1031. [Google Scholar] [CrossRef]
  29. Gopalakrishnan, S.; Zampieri, A.; Schwieger, W. Mesoporous ZSM-5 zeolites via alkali treatment for the direct hydroxylation of benzene to phenol with N2O. J. Catal. 2008, 260, 193–197. [Google Scholar] [CrossRef]
  30. Su, L.L.; Zhang, J.Q.; Wang, H.X.; Li, Y.G.; Shen, W.J.; Xu, Y.D.; Bao, X.H. Creating mesopores in ZSM-5 zeolite by alkali treatment: A new way to enhance the catalytic performance of methane dehydroaromatization on Mo/HZSM-5 catalysts. Catal. Lett. 2003, 91, 155–167. [Google Scholar] [CrossRef]
  31. Xiang, Q.J.; Lv, K.L.; Yu, J.G. Pivotal role of fluorine in enhanced photocatalytic activity of anatase TiO2 nanosheets with dominant (001) facets for the photocatalytic degradation of acetone in air. Appl. Catal. B Environ. 2010, 96, 557–564. [Google Scholar] [CrossRef]
  32. Hou, C.; Liu, W.; Zhu, J. Synthesis of NaOH-Modified TiOF2 and Its Enhanced Visible Light Photocatalytic Performance on RhB. Catalysts 2017, 7, 243. [Google Scholar] [CrossRef]
  33. Yang, H.G.; Zeng, H.C. Preparation of hollow anatase TiO2 nanospheres via Ostwald ripening. Phys. Chem. B 2004, 108, 3492–3495. [Google Scholar] [CrossRef]
  34. Zhao, W.; Liu, N.; Wang, H.; Mao, L. Sacrificial template synthesis of coreshell SrTiO3/TiO2heterostructured microspheres photocatalyst. Ceram. Int. 2017, 43, 4807–4813. [Google Scholar] [CrossRef]
  35. Zo, M.; Liu, H.; Feng, L.; Xiong, F.; Thoma, T.; Yang, M. Effect of nitridation on visible light photocatalytic behavior of microporous (Ag, Ag2O) co-loaded TiO2. Microporous Mesoporous Mater. 2017, 240, 137–144. [Google Scholar] [CrossRef]
  36. Guo, S.-Y.; Dai, J.-G.; Zhao, T.-J.; Hou, S.-D.; Zhang, P.; Wang, P.; Sun, G.-X. A novel microporous morphous-ZnO@TiO2/graphene ternary nanocomposite with enhanced photocatalytic activity. RSC Adv. 2017, 7, 36787–36792. [Google Scholar] [CrossRef]
  37. Tayade, R.J.; Surolia, P.K.; Kulkarni, R.G.; Jasra, R.V. Photocatalytic degradation of dyes and organic contaminants in water using nanocrystalline anatase and rutile TiO2. Sci. Technol. Adv. Mater. 2007, 8, 455–462. [Google Scholar] [CrossRef]
  38. Xu, S.H.; Shangguan, W.F.; Yuan, J. Synthesis and performance of novel magnetically separable nanospheres of titanium dioxide photocatalyst with egg-like structure. Nanotechnology 2008, 19. [Google Scholar] [CrossRef] [PubMed]
  39. Iwata, T.; Watanabe, A.; Iseki, M.; Watanabe, M.; Kandor, H. Strong Donation of the Hydrogen Bond of Tyrosine during Photoactivation of the BLUF Domain. Phys. Chem. Lett. 2011, 2, 1015–1019. [Google Scholar] [CrossRef]
  40. Ning, Y. Structural Identification of Organic Compounds and Organic Spectroscopy; Science Press: Beijing, China, 2000; ISBN 978-7-03-0074-126. [Google Scholar]
  41. Li, D.; Haneda, H.; Hishita, S.; Ohashi, N. Visible-Light-Driven N-F-Codoped TiO2 Photocatalysts. 2. Optical Characterization, Photocatalysis, and Potential Application to Air Purification. Chem. Mater. 2005, 17, 2596–2602. [Google Scholar] [CrossRef]
  42. Zhang, J.; Hou, X.; Pang, Z.; Cai, Y.; Zhou, H.; Lv, P.; Wei, Q. Fabrication of hierarchical TiO2 nanofibers by microemulsion electrospinning for photocatalysis applications. Ceram. Int. 2017, 43, 15911–15917. [Google Scholar] [CrossRef]
  43. Hu, Y.; Li, D.; Zheng, Y.; Chen, W.; He, Y.; Shao, Y.; Fu, X.; Xiao, G. BiVO4/TiO2 nanocrystalline heterostructure: A wide spectrum responsive photocatalyst towards the highly efficient decomposition of gaseous benzene. Appl. Catal. B Environ. 2011, 104, 30–36. [Google Scholar] [CrossRef]
  44. Park, H.; Choi, W. Photocatalytic Reactivities of Nafion-Coated TiO2 for the Degradation of Charged Organic Compounds under UV or Visible Light. J. Phys. Chem. B 2005, 109, 11667–11674. [Google Scholar] [CrossRef] [PubMed]
Figure 1. X-ray diffraction (XRD) patterns of as-prepared F-TiO2 and OH-TiO2 samples.
Figure 1. X-ray diffraction (XRD) patterns of as-prepared F-TiO2 and OH-TiO2 samples.
Catalysts 08 00070 g001
Figure 2. Field emission Scanning electron microscopy (FE-SEM) and high-resolution transmission electron microscope (HRTEM) of as-prepared F-TiO2 and OH-TiO2 samples. (a) FE-SEM images of F-TiO2; (b) FE-SEM images of OH-TiO2; (c,d) HRTEM images of F-TiO2; (e,f) HRTEM images of OH-TiO2.
Figure 2. Field emission Scanning electron microscopy (FE-SEM) and high-resolution transmission electron microscope (HRTEM) of as-prepared F-TiO2 and OH-TiO2 samples. (a) FE-SEM images of F-TiO2; (b) FE-SEM images of OH-TiO2; (c,d) HRTEM images of F-TiO2; (e,f) HRTEM images of OH-TiO2.
Catalysts 08 00070 g002
Figure 3. Raman spectroscopy of catalysts.
Figure 3. Raman spectroscopy of catalysts.
Catalysts 08 00070 g003
Figure 4. (a) UV-Vis diffuse reflectance spectra and (b) bandgap of as-prepared F-TiO2 and OH-TiO2 samples.
Figure 4. (a) UV-Vis diffuse reflectance spectra and (b) bandgap of as-prepared F-TiO2 and OH-TiO2 samples.
Catalysts 08 00070 g004
Figure 5. (a) Fourier transform infrared (FT-IR) spectra and (b) Photoluminescence (PL) emission spectra of as-prepared F-TiO2 and OH-TiO2 samples.
Figure 5. (a) Fourier transform infrared (FT-IR) spectra and (b) Photoluminescence (PL) emission spectra of as-prepared F-TiO2 and OH-TiO2 samples.
Catalysts 08 00070 g005
Figure 6. Photocatalytic performance of OH-TiO2 and F-TiO2 for MB and RhB and UV-vis absorption spectral changes during the photocatalytic degradation of MB (a) photocatalytic performance of MB and RhB samples; (b) photocatalytic performance of OH-TiO2 with varying amounts of catalysts; (c) UV-vis absorption spectral changes during the photocatalytic degradation of MB with OH-TiO2; (d) UV-vis absorption spectral changes during the photocatalytic degradation of MB with F-TiO2.
Figure 6. Photocatalytic performance of OH-TiO2 and F-TiO2 for MB and RhB and UV-vis absorption spectral changes during the photocatalytic degradation of MB (a) photocatalytic performance of MB and RhB samples; (b) photocatalytic performance of OH-TiO2 with varying amounts of catalysts; (c) UV-vis absorption spectral changes during the photocatalytic degradation of MB with OH-TiO2; (d) UV-vis absorption spectral changes during the photocatalytic degradation of MB with F-TiO2.
Catalysts 08 00070 g006
Figure 7. Radical-scavenging experiments of OH-TiO2 and F-TiO2 toward MB.
Figure 7. Radical-scavenging experiments of OH-TiO2 and F-TiO2 toward MB.
Catalysts 08 00070 g007

Share and Cite

MDPI and ACS Style

Hou, C.; Zhu, J.; Song, Q. Low Temperature Synthesis of Nest-Like Microsphere with Exposed (001) Facets and Its Enhanced Photocatalytic Performance by NaOH Alkalization. Catalysts 2018, 8, 70. https://doi.org/10.3390/catal8020070

AMA Style

Hou C, Zhu J, Song Q. Low Temperature Synthesis of Nest-Like Microsphere with Exposed (001) Facets and Its Enhanced Photocatalytic Performance by NaOH Alkalization. Catalysts. 2018; 8(2):70. https://doi.org/10.3390/catal8020070

Chicago/Turabian Style

Hou, Chentao, Jiaming Zhu, and Qiaoqiao Song. 2018. "Low Temperature Synthesis of Nest-Like Microsphere with Exposed (001) Facets and Its Enhanced Photocatalytic Performance by NaOH Alkalization" Catalysts 8, no. 2: 70. https://doi.org/10.3390/catal8020070

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