Next Article in Journal
Galactooligosaccharide Production from Pantoea anthophila Strains Isolated from “Tejuino”, a Mexican Traditional Fermented Beverage
Next Article in Special Issue
Hydrogen Production from Cyclic Chemical Looping Steam Methane Reforming over Yttrium Promoted Ni/SBA-16 Oxygen Carrier
Previous Article in Journal
Improved H2 Production by Ethanol Steam Reforming over Sc2O3-Doped Co-ZnO Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of NaOH-Modified TiOF2 and Its Enhanced Visible Light Photocatalytic Performance on RhB

College of Geology and Environment, Xi’an University of Science and Technology, Xi’an 710054, China
*
Author to whom correspondence should be addressed.
Catalysts 2017, 7(8), 243; https://doi.org/10.3390/catal7080243
Submission received: 16 July 2017 / Revised: 7 August 2017 / Accepted: 16 August 2017 / Published: 22 August 2017
(This article belongs to the Special Issue Nanomaterials for Environmental Purification and Energy Conversion)

Abstract

:
NaOH-modified TiOF2 was successfully prepared using a modified low-temperature hydrothermal method. Scanning electron microscopy shows that NaOH-modified TiOF2 displayed a complex network shape with network units of about 100 nm. The structures of NaOH-modified TiOF2 have not been reported elsewhere. The network shape permits the NaOH-modified TiOF2 a SBET of 36 m2∙g−1 and a pore diameter around 49 nm. X-ray diffraction characterization shows that TiOF2 and NaOH-modified TiOF2 are crystallized with a pure changed cubic phase which accords with the SEM results. Fourier transform infrared spectroscopy characterization shows that NaOH-modified TiOF2 has more O–H groups to supply more lone electron pairs to transfer from O of O–H to Ti and O of TiOF2. UV–vis diffuse reflectance spectroscopy (DRS) shows that the NaOH-modified TiOF2 sample has an adsorption plateau rising from 400 to 600 nm in comparison with TiOF2, and its band gap is 2.62 eV, lower than that of TiOF2. Due to the lower band gap, more O–H groups adsorption, network morphologies with larger surface area, and sensitization progress, the NaOH-modified TiOF2 exhibited much higher photocatalytic activity for Rhodamine B (RhB) degradation. In addition, considering the sensitization progress, O–H groups on TiOF2 not only accelerated the degradation rate of RhB, but also changed its degradation path. As a result, the NaOH-modified TiOF2 exhibited much higher photocatalytic activity for RhB degradation than the TiOF2 in references under visible light. This finding provides a new idea to enhance the photocatalytic performance by NaOH modification of the surface of TiOF2.

Graphical Abstract

1. Introduction

Nowadays, environmental pollution is affecting human survival and development. Photocatalysis is considered an efficient, stable, and environmentally friendly method for controlling environmental pollution [1]. In the past, TiO2 has been widely used as a photocatalyst in the photo-degradation of organic pollutants. However, it has a wide energy band gap (3.1–3.2 eV) which only permit its UV light response and can easily cause electron–hole recombination [1,2,3,4]. Thus, studies on changing morphology [1,2,3], modification [1,4,5], and other methods were conducted to decrease its band gap or inhibit its electron–hole recombination. The discovery of non-titanium semiconductor photocatalysts with a narrow intrinsic energy band gap, efficiently driven by visible light, may also attract much attention [5,6,7,8,9,10,11,12,13].
Recently, Li’s research group found that TiOF2 cubes—considered a promising anode material for lithium ion batteries (LIBs) [14,15,16,17,18,19]—showed visible-light driven property and exhibited excellent performance in photodegradation of Rhodamine B (RhB) and 4-chlorophenol (4-CP) [7]. TiOF2 is also proven to be more active and durable at room temperature due to the covalent bonds of F species with Ti [8,9]. Only a few studies focused on the photocatalytic activity of TiOF2 have been reported [7,8,9]. As usual, TiOF2 nanoparticles were synthesized via hydrothermal [10,14,15,16,17,18] and solvothermal [7,11,12] methods from titanium (IV) isopropoxide (TIP), and have a cubic shape [7,8,9,10,11,12,13,14,15,16,17,18]. The size of TiOF2 nanocubes could be affected by alcoholysis time, alcohol kind, solvothermal temperature, different H2O production rate and amount [7,17]. While the photocatalytic activity of TiOF2 is still unsatisfactory, it is necessary to explore novel approaches to improve its photocatalytic performance.
Alkali modification is proven to be an effective method to enhance the catalytic performance for α-pinene isomerization, formaldehyde oxidation, and benzene hydroxylation [20,21,22,23]. Thus, it stimulated us to modify the TiOF2 catalyst obtained from our earlier studies. In this study, we firstly reported a network-shaped NaOH-modified TiOF2 treated by a hydrothermal process under low temperature. The FTIR measurement showed that more associated O–H exists on the surface of TiOF2, which can remarkably enhance the catalytic activity of TiOF2 toward RhB oxidation under visible light. The NaOH-modified TiOF2 had better photocatalytic performance than TiOF2 in Li’s research [7].

2. Results and Discussion

2.1. Phase Structures and Morphology

The phase and crystallinity of the as-prepared TiOF2 and NaOH-modified TiOF2 samples were tested by XRD analysis. It can be seen from Figure 1, the patterns of as-prepared TiOF2 and NaOH-modified TiOF2 samples all have sharp peaks at 2θ = 23.6°, 48.1°, and 54.2°, corresponding to the (100), (200), and (210) planes of the cubic TiOF2 phase (JCPDS no. 08-0060) [7,13] and no peak of any anatase TiO2 (JCPDS no. 21-1272) [24,25] crystal appears, indicating that the as-prepared samples have high crystallinity and pure phase of cubic TiOF2. It also indicates that the height of the (100) crystal planes of the TiOF2 and NaOH-modified TiOF2 are higher, and the (200) and (210) crystal planes are lower than that of the standard cubic TiOF2, indicating a new shape of the as-prepared TiOF2 and NaOH-modified TiOF2. The Scherrer formula was used to calculate the normal distance of certain crystal surfaces of TiOF2 and NaOH-modified TiOF2
τ = κ λ / ( β cos θ )
where τ , κ , λ , β , and θ are the mean normal distance of certain crystal surfaces, the shape constant with a value of 0.89 when β is the half width of the diffraction peak (FWHM), the diffracted ray wavelength (0.15418 nm for Cu-Ka), and the diffraction angle in radians, respectively [26].
The normal distances of TiOF2 are 7.26, 17.81, 11.67, 10.77, 17.38, and 16.98 nm along the (100), (110), (200), (210), (220), and (215) planes [7,13], respectively, while the distanced of crystal NaOH-modified TiOF2 change to 15.99, 42.18, 26.26, 30.78, 40.80, and 48.39 nm along the corresponding planes, respectively. It can be seen that the normal distance of the crystalline phase of NaOH-modified TiOF2 shrunk 8.88% and 3.72% along (100) and (110) planes. However, an increase was observed along (200), (210), and (215) planes. This can be explained in that NaOH-modifying induces more O–H adsorbed onto (100) and (110) planes of TiOF2 and, thus, induces the planes’ exposure.
Morphologies and microstructure of original TiOF2 and NaOH-modified TiOF2 were checked by SEM characterization. In Figure 2a,b, TiOF2 crystals displayed a mixture of the cubic image which is accords with the cubic image in [7,8,9,10,11,12,13,14,15,16,17,18]. Each individual particle crystal is about 50–300 nm and tends to aggregate, forming larger particles while—in Figure 2c,d—the NaOH-modified TiOF2 displayed a more complex network shape with network units in about 100 nm. The NaOH-modified TiOF2 shows phases assembling along certain directions. This accords with the XRD results. The network shape permits much more surface area for photocatalysis. These structures of NaOH-modified TiOF2 have not been reported elsewhere. The Barrett-Joyner-Halenda (BJH) method was used to analyze the pore size distribution and pore volume and the surface area (SBET) was calculated using the BET method. The Figure 3a demonstrated that the NaOH-modified TiOF2 showed a typical IV type N2 adsorption–desorption isotherm and mesoporous structure with an average pore diameter of about 49 nm. Thus, its SBET can reach as high as 36 m2∙g−1, while the average pore diameter and SBET of TiOF2 are only 3 nm and 2.7 m2∙g−1, which is much lower than that of NaOH-modified TiOF2. The larger surface area permits more O–H and pollutant adsorption and the formation of additional mesopores affects the improvement of mass transfer, enhancing photocatalytic performance accordingly [21,22].

2.2. FTIR Analysis

Figure 3b shows the FTIR spectra of TiOF2 and NaOH-modified TiOF2. The strong band around 700–500 cm−1 could contribute to the Ti–O–Ti stretching vibration [22,23,24]. The peak around 3379 cm−1 and the broad band centered around 3212 cm−1 were due to the free and bonding O–H stretching vibration of Ti−OH, respectively [27,28]. The peak at 1620 cm−1 was due to the O–H bending vibration of Ti−OH [16,22,29,30,31,32,33]. The broad band centered around 3212 cm−1 in NaOH-modified TiOF2 becomes broader than that in TiOF2, meaning that more O–H bonds or associated O–H appeared in NaOH-modified TiOF2. According to previous work, the free O–H stretching vibration used to appear at about 3600 cm−1 without bonding O–H [27,28]. It can be seen that the O–H frequency for TiOF2 and NaOH-modified TiOF2 is 221 cm−1 and 388 cm−1 lower than 3600 cm−1, indicating a strong hydrogen bond impact [27,28]. The O–H on the TiO2 surface can enhance the transference of photo-generated electrons and then enhance photocatalytic performance [29]. The peaks around 930 cm−1 were due to the Ti–F vibrations in the TiOF2 [16]. The peak intensity decreased from TiOF2 to NaOH-modified TiOF2, indicating F was exchanged by O–H after NaOH modification. All of these show that the NaOH-modified TiOF2 samples contain more O–H groups than TiOF2. It can be explained that TiOF2 was modified in NaOH solution, thus, more O–H would be chemisorbed onto TiOF2, and further exchanged with F. Then, more lone pair electrons in the O–H groups transferring from the O of O–H to Ti and the O of TiOF2, the performance of TiOF2 can be enhanced accordingly [21]. In addition, because RhB is a cationic dye, NaOH brings more O–H onto the surface of TiOF2 to hold more RhB and accelerate its degradation rate [34,35].

2.3. UV–Vis Analysis

Figure 4 shows that the UV–vis absorption spectroscopy and band gap of as-prepared TiOF2 and NaOH-modified TiOF2 samples. The NaOH-modified TiOF2 has a raised adsorption plateau from 400 to 600 nm, which indicates stronger visible light absorption than that of TiOF2 (Figure 4a). Band gap estimation can be seen in Figure 4b showing that the band gap of NaOH-modified TiOF2 is 2.62 eV, which is lower than that of TiOF2 (2.80 eV) and lower than anatase TiO2 (3.2 eV) [1,2,3,4], NiO (4.0 eV) [6], and other oxides, indicating easier excitation by visible light. This can be explained in that NaOH treatment causes certain facet exposure and network morphologies of TiOF2, changing its light absorption properties. Thus, the NaOH treatment lowered the band gap of TiOF2, enhanced its visible light absorption, and further enhanced its visible light photocatalytic properties.

2.4. Catalytic Activity

Figure 5 shows the visible light photcatalytic properties of TiOF2, NaOH-modified TiOF2, TiOF2 in reference (TiOF2-Ref) [7,13], TiOF2-crushed in reference (TiOF2-crushed-Ref) [7], and P25. It can be seen in Figure 5a that, in the adsorption test in dark and in light on the process without the catalyst for RhB, the decrease of RhB is very small. It can be concluded that the adsorption and sensitization mechanisms can be negligible in the degradation process. Thus, the degradation of RhB was a photocatalytic process. The concentration of RhB decreased under the same conditions, which means that all samples are visible-light active. It also shows that NaOH-modified TiOF2 can cause almost complete decomposition of RhB in 3 h, having better photocatalytic performance than all of the TiOF2 in reference [7,13]. While P25 and TiOF2 performed poorly compared to NaOH-modified TiOF2 and TiOF2-Ref. The reaction rate of all of the samples are shown in Figure 5b. It can be seen that the data was fitted with the first-order reaction equation as
ln ( C 0 / C ) = k t
where t is the reaction time, C 0 is concentration of RhB at time 0, C is the concentration of RhB at time t , and k is the reaction rate constant.
It can be seen that P25 and TiOF2 had rate constants of only 0.10 and 0.16 h−1, indicating poor photocatalytic performance. The result is consistent with previous work [8,20]. The calculated rate constants are 1.37, 0.73, and 1.24 h−1 for NaOH-modified TiOF2, TiOF2-Ref, and TiOF2-crushed-Ref, respectively. The NaOH-modified TiOF2 sample shows the best performance among all the photocatalysts, whose degradation rates are much higher than that of P25 and TiOF2 in our samples and 10.4% higher than that of TiOF2-crushed-Ref The excellent performance could be mainly attributed to its larger SBET (32 m2∙g−1 for TiOF2-crushed-Ref) and more bonding O–H [7,21].

2.5. Effect of the Sensitization Mechanism

According to previous studies, dyes can be degraded on TiO2 through a sensitized process under visible light [34,35]. In order to know whether there is a similar path on TiOF2 and NaOH-modified TiOF2, UV–vis absorption spectral changes of RhB with visible light irradiation time in the suspension of TiOF2 and NaOH-modified TiOF2 were tested. The results are shown in Figure 6. It can be seen in Figure 6 that the spectral change of RhB with the irradiation time on NaOH-modified TiOF2 is quite different from that on TiOF2. There is a blue-shift from 558 to 498 nm in the absorption maximum with irradiation time for NaOH-modified TiOF2, while none in that of TiOF2. This is attributed to the N-deethylation products of RhB, which confirms the possibility of the sensitization mechanism [34,35]. Thus, NaOH treatment can induce the sensitization process and change the degradation path of TiOF2.

3. Materials and Methods

Tetrabutyl titanate (TBOT, A.R. grade) was purchased from Fu Chen Chemical Reagent Factory, Tianjin, China. Absolute ethyl alcohol (C2H5OH, A.R. grade) and sodium hydroxide (NaOH, A.R. grade) was purchased from Fuyu Fine Chemical Co., Ltd., Tianjin, China. Hydrofluoric acid (HF, A.R. grade) was purchased from Xilong Chemical Industry Co., Ltd., Chengdu, China. All reagents are used without further purification. Ultrapure water was used as the experimental water.
NaOH-modified TiOF2 was synthesized via a modified low-temperature hydrothermal method. In a typical synthesis, 30.4 mL absolute ethyl alcohol was added into 35.2 mL TBOT, which was named solution A. Absolute ethyl alcohol (30.4 mL) and 20.2 mL HF were added into 180 mL ultrapure water, which was named solution B. Solution A was dropped into solution B under medium-speed magnetic stirring at 20 °C for 1.5 h to obtain a faint yellow sol. The sol was aged at room temperature for 2 days to change to a gel. The gel was then transferred into a 50-mL Teflon-lined stainless steel autoclave. When sealed, the autoclave was placed at 100 °C for 2 h in a drying box, then was naturally cooled to room temperature. Ultra-pure water and absolute ethanol were used to wash the obtained white precipitates several times to reach a pH of 7, and then the precipitates were dried at 100 °C. The as-prepared sample was TiOF2. One gram of the TiOF2 precursor was dispersed in 100 mL 5 mol·L−1 NaOH solution under magnetic stirring with a speed of 4000 r·min−1 for 1 h, then the suspension was also washed with ultra-pure water and absolute ethanol to reach a pH of 7. The product was dried at 100 °C for 12 h. The sample was denoted as NaOH-modified TiOF2. The crystal structure as analyzed by a XD-2 X-ray diffractometer (Beijing Purkinje, Beijing, China) with Cu Kαradiation with a scan rate of 4.0000°·min−1. The morphology was examined by field emission scanning electron microscopy (FESEM, JEOL JSM6700, Tokyo, Japan). Fourier transform infrared (FTIR) spectra were recorded using a Bruker TENSOR27 (Karlsurhe, Germeny) using the KBr method. The optical properties were determined by UV–vis diffuse reflectance spectroscopy (UV–vis DRS: (Shimadzu 2600, Beijing, China). N2 adsorption-desorption isotherms were measured at 77 K and the BET method was used to calculate the surface area (SBET) by a JW-BK122F (Beijing, China).
The degradation of RhB was conducted at room temperature in a 150 mL double-layered quartz reactor containing 50 mg catalyst and 50 mL 5.0 mg·L−1 RhB solution. A 300 W Xe lamp (Jiguang-300, Shanghai, China) was located at a distance of 15 cm from the RhB solution to simulate solar light. A cutoff filter (JB-420, Shanghai, China) was chosen to filter off the light whose wavelength was less than 420 nm to simulate visible light. The solution was magnetically stirred for 30 min to ensure the adsorption–desorption equilibrium, then the xenon lamp was turned on to start the photocatalytic degradation. At 30 min time intervals, about 5.0 mL RhB solution was extracted and centrifuged at high-speed (11,000 r·min−1) to remove catalysts. Then the concentration of the remaining RhB solutions were analyzed with a Purkinje UV1901 UV–vis spectrophotometer at 554 nm. The photocatalyst was separated from the RhB solution and another run of the reaction was started to investigate the durability of the catalysts.

4. Conclusions

NaOH-modified TiOF2 was successfully prepared via a modified low-temperature solvothermal method. It exhibited much better photocatalytic performance for RhB degradation. XRD characterization shows that TiOF2 and NaOH-modified TiOF2 are crystallized with a pure changed cubic phase which is accord with the SEM results. SEM shows that TiOF2 crystals displayed a mixture of the cubic images, while the NaOH-modified TiOF2 displayed a more complex network shape with network units in about 100 nm. These structures of NaOH-modified TiOF2 have not been reported elsewhere. The network shape permits the NaOH-modified TiOF2 a surface area of 36 m2·g−1 and a pore diameter about 49 nm, which will enhance the adsorption of O–H groups and pollutants. FTIR characterization shows that NaOH-modified TiOF2 has more O–H groups to supply more lone electron pairs transferring from O of the O–H groups to Ti and O of TiOF2, in accordance with the BET analysis. UV–vis absorption spectroscopy shows that the NaOH-modified TiOF2 samples have an adsorption plateau rising from 400 to 600 nm in comparison with TiOF2 and its band gap is 2.62 eV, lower than that of TiOF2. Due to the lower band gap, more O–H groups adsorption, network morphologies with larger surface area, and sensitization process, the NaOH-modified TiOF2 exhibited much higher photocatalytic activity for RhB degradation. In addition, considering the sensitization process, O–H on TiOF2 not only accelerated the degradation rate of RhB, but also changed its degradation path. This finding provides a new idea to enhance the photocatalytic performence by NaOH modification of the surface of TiOF2. Considering its synthesizing process, the NaOH-modified TiOF2 needs much lower temperature and shorter time than TiOF2-crushed-Ref, but has much better photocatalytic performance, which provides a more economic choice.

Acknowledgments

Financial support was provided by Shaanxi Key Industrial Projects (2014GY2-07) and the Shaanxi Province Education Department Science and Technology Research Plan (15JK1460).

Author Contributions

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

Conflicts of Interest

There is no conflict of interest existing in the manuscript submission, and it is approved by all of the authors for publication. All the authors listed have approved the manuscript to be enclosed.

References

  1. Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications, and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  2. Nunes, D.; Pimentel, A.; Santos, L.; Barquinha, P.; Fortunato, E.; Martins, R. Photocatalytic TiO2 Nanorod Spheres and Arrays Compatible with Flexible Applications. Catalysts 2017, 7, 60. [Google Scholar] [CrossRef]
  3. Zuo, S.; Jiang, Z.; Liu, W.; Yao, C.; Chen, Q.; Liu, X. Synthesis and Characterization of Urchin-like Mischcrystal TiO2 and Its Photocatalysis. Mater. Charact. 2014, 96, 177–182. [Google Scholar] [CrossRef]
  4. Xie, C.; Yang, S.; Shi, J.; Niu, C. Highly Crystallized C-Doped MesoporousAnatase TiO2 with Visible Light Photocatalytic Activity. Catalysts 2016, 6, 117. [Google Scholar] [CrossRef]
  5. Cherian, C.T.; Reddy, M.V.; Magdaleno, T.; Sow, C.-H.; Ramanujachary, K.V.; Rao, G.V.S.; Chowdari, B.V.R. (N,F)-Co-doped TiO2: Synthesis, anatase-rutile conversion and Li-cycling properties. CrystEngComm 2012, 14, 978–986. [Google Scholar] [CrossRef]
  6. Ramasami, A.K.; Reddy, M.V.; Balakrishna, G.R. Combustion synthesis and characterization of NiO nanoparticles. Mater. Sci. Semicond. Process. 2015, 40, 194–202. [Google Scholar] [CrossRef]
  7. Wang, J.; Cao, F.; Bian, Z.; Leung, M.K.H.; Li, H. Ultrafine single-crystal TiOF2 nanocubes with mesoporous structure, high activity and durability in visible light driven photocatalysis. Nanoscale 2014, 6, 897–902. [Google Scholar] [CrossRef] [PubMed]
  8. Wang, Z.; Ana, C.; Zhang, J. Synthesis of heterostructured Pd@TiO2/TiOF2 nanohybrids with enhanced photocatalytic performance. Mater. Res. Bull. 2016, 80, 337–343. [Google Scholar]
  9. Dongn, P.; Cui, E.; Hou, G.; Guan, R.; Zhang, Q. Synthesis and photocatalytic activity of Ag3PO4/TiOF2 composites with enhanced stability. Mater. Lett. 2015, 143, 20–23. [Google Scholar] [CrossRef]
  10. Lv, K.; Yu, J.; Cui, L.; Chen, S.; Li, M. Preparation of thermally stable anatase TiO2 photocatalyst from TiOF2 precursor and its photocatalytic activity. J. Alloys Compd. 2011, 509, 4557–4562. [Google Scholar] [CrossRef]
  11. Wang, Z.; Huang, B.; Dai, Y.; Zhang, X.; Qin, X.; Li, Z.; Zheng, Z.; Cheng, H.; Guo, L. Topotactic transformation of single-crystalline TiOF2 nanocubes to ordered arranged 3D hierarchical TiO2 nanoboxes. CrystEngComm 2012, 14, 4578–4581. [Google Scholar] [CrossRef]
  12. Huang, Z.; Wang, Z.; Lv, K.; Zheng, Y.; Deng, K. Transformation of TiOF2 cube to a hollow nanobox assembly from anatase TiO2 nanosheets with exposed {001} facets via solvothermal strategy. ACS Appl. Mater. Interfaces 2013, 5, 8663–8669. [Google Scholar] [CrossRef] [PubMed]
  13. Wen, C.Z.; Hu, Q.H.; Guo, Y.N.; Gong, X.Q.; Qiao, S.Z.; Yang, H.G. From titanium oxydifluoride (TiOF2) to titania (TiO2): Phase transition and non-metal doping with enhanced photocatalytic hydrogen (H2) evolution properties. Chem. Commun. 2011, 47, 6138–6140. [Google Scholar] [CrossRef] [PubMed]
  14. Jung, M.; Kim, Y.; Lee, Y. Enhancement of the electrochemical capacitance of TiOF2 obtained via control of the crystal structure. J. Ind. Eng. Chem. 2017, 47, 187–193. [Google Scholar] [CrossRef]
  15. Louvain, N.; Karkar, Z.; El-Ghozzi, M.; Bonnet, P.; Gúerin, K.; Willmann, P. Fluorinationof anatase TiO2 towards titanium oxyfluoride TiOF2: A novel synthesis approach and proof of the Li-insertion mechanism. J. Mater. Chem. A 2014, 2, 15308–15315. [Google Scholar] [CrossRef]
  16. Li, W.; Body, M.; Legein, C.; Dambournet, D. Identify OH groups in TiOF2 and their impact on the lithium intercalation properties. J. Solid State Chem. 2017, 246, 113–118. [Google Scholar] [CrossRef]
  17. Chen, L.; Shen, L.; Nie, P.; Zhang, X.; Li, H. Facile hydrothermal synthesis of single crystalline TiOF2 nanocubes and their phase transitions to TiO2 hollow nanocages as anode materials for lithium-ion battery. Electrochim. Acta 2012, 62, 408–415. [Google Scholar] [CrossRef]
  18. Reddy, M.V.; Madhavi, S.; Rao, G.V.S.; Chowdari, B.V.R. Metal oxyfluorides TiOF2 and NbO2F as anodes for Li-ion batteries. J. Power Sources 2006, 162, 1312–1321. [Google Scholar] [CrossRef]
  19. Reddy, M.V.; Rao, G.V.S.; Chowdari, B.V.R. Metal Oxides and Oxysalts as Anode Materials for Li Ion Batteries. Chem. Rev. 2013, 113, 5364–5457. [Google Scholar] [CrossRef] [PubMed]
  20. Zhou, P.; Yu, J.; Nie, L.; Jaroniec, M. Dual-dehydrogenation-promoted catalytic oxidation of formaldehyde on alkali-treated Pt clusters at room temperature. J. Mater. Chem. A 2015, 3, 10432–10438. [Google Scholar] [CrossRef]
  21. Nie, L.; Yu, J.; Li, X.; Cheng, B.; Liu, G.; Jaroniec, M. Enhanced Performance of NaOH-Modified Pt/TiO2 toward Room Temperature Selective Oxidation of Formaldehyde. Environ. Sci. Technol. 2013, 47, 2777–2783. [Google Scholar] [CrossRef] [PubMed]
  22. 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]
  23. Wu, Y.; Tian, F.; Liu, J.; Song, D.; Jia, C.; Chen, Y. Enhanced catalytic isomerization of α-pinene over mesoporous zeolite beta of low Si/Al ratio by NaOH treatment. Microporous Mesoporous Mater. 2012, 162, 168–174. [Google Scholar] [CrossRef]
  24. Wang, Z.; Lv, K.; Wang, G.; Deng, K.; Tang, D. Study on the shape control and photocatalytic activity of high-energy anatasetitania. Appl. Catal. B Environ. 2010, 100, 378–385. [Google Scholar] [CrossRef]
  25. Niu, L.; Zhang, Q.; Liu, J.; Qian, J.; Zhou, X. TiO2 nanoparticles embedded in hollow cube with highly exposed {001} facets: Facile synthesis and photovoltaic applications. J. Alloys Compd. 2016, 656, 863–870. [Google Scholar] [CrossRef]
  26. He, M.; Wang, Z.; Yan, X.; Tian, L.; Liu, G.; Chen, X. Hydrogenation effects on the lithium ion battery performance of TiOF2. J. Power Sources 2016, 306, 309–316. [Google Scholar] [CrossRef]
  27. Iwata, T.; Watanabe, A.; Iseki, M.; Watanabe, M.; Kandor, H. Strong Donation of the Hydrogen Bond of Tyrosine during Photoactivation of the BLUF Domain. J. Phys. Chem. Lett. 2011, 2, 1015–1019. [Google Scholar] [CrossRef]
  28. Ning, Y. Structural Identification of Organic Compounds and Organic Spectroscopy; Science Press: Beijing, China, 2000; ISBN 978-7-03-0074-126. [Google Scholar]
  29. Zhang, Y.; Zhang, Q.; Xia, T.; Zhu, D.; Chen, Y.; Chen, X. The influence of reaction temperature on the formation and photocatalytic hydrogen generation of (001) faceted TiO2 nanosheets. ChemNanoMat 2015, 1, 270–275. [Google Scholar] [CrossRef]
  30. Vaiano, V.; Sannino, D.; Almeida, A.R.; Mul, G.; Ciambelli, P. Investigation of the Deactivation Phenomena Occurring in the Cyclohexane Photocatalytic Oxidative Dehydrogenation on MoOx/TiO2 through Gas Phase and in situ DRIFTS Analyses. Catalysts 2013, 3, 978–997. [Google Scholar] [CrossRef]
  31. Samsudin, E.M.; Hamid, S.B.A.; Basirun, W.J.; Centi, G. Synergetic effects in novel hydrogenated F-doped TiO2 photocatalysts. Appl. Surf. Sci. 2016, 370, 380–393. [Google Scholar] [CrossRef]
  32. Ding, X.; Hong, Z.; Wang, Y.; Lai, R.; Wei, M. Synthesis of square-like anatase TiO2 nanocrystals based on TiOF2 quantum dots. J. Alloys Compd. 2013, 550, 475–478. [Google Scholar] [CrossRef]
  33. Nishikiori, H.; Hayashibe, M.; Fujii, T. Visible Light-Photocatalytic Activity of Sulfate-Doped Titanium Dioxide Prepared by the Sol−Gel Method. Catalysts 2013, 3, 363–377. [Google Scholar] [CrossRef]
  34. Kowalska, E.; Remita, H.; Colbeau-Justin, C.; Hupka, J.; Belloni, J. Modification of Titanium Dioxide with Platinum Ions and Clusters: Application in Photocatalysis. J. Phys. Chem. C 2008, 112, 1124–1131. [Google Scholar] [CrossRef]
  35. 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. XRD pattern with 2θ from 10° to 80° (a) and partial enlarged detail of 2θ from 47° to 56° (b) of as-prepared TiOF2, NaOH-modified TiOF2, commercial P25 and a standard card of TiOF2 and TiO2.
Figure 1. XRD pattern with 2θ from 10° to 80° (a) and partial enlarged detail of 2θ from 47° to 56° (b) of as-prepared TiOF2, NaOH-modified TiOF2, commercial P25 and a standard card of TiOF2 and TiO2.
Catalysts 07 00243 g001
Figure 2. SEM of as-synthesized samples: (a,b) TiOF2; and (c,d) NaOH-modified TiOF2.
Figure 2. SEM of as-synthesized samples: (a,b) TiOF2; and (c,d) NaOH-modified TiOF2.
Catalysts 07 00243 g002
Figure 3. N2 adsorption-desorption isotherm of the NaOH-modified TiOF2 (a) and the FTIR spectra for TiOF2 and NaOH-modified TiOF (b).
Figure 3. N2 adsorption-desorption isotherm of the NaOH-modified TiOF2 (a) and the FTIR spectra for TiOF2 and NaOH-modified TiOF (b).
Catalysts 07 00243 g003
Figure 4. UV–vis DRS spectra (a) and band gap (b) of TiOF2 and NaOH-modified TiOF2.
Figure 4. UV–vis DRS spectra (a) and band gap (b) of TiOF2 and NaOH-modified TiOF2.
Catalysts 07 00243 g004
Figure 5. Catalytic activity of RhB under visible light: (a) concentration dependent on time and (b) kinetic fit for the degradation of RhB.
Figure 5. Catalytic activity of RhB under visible light: (a) concentration dependent on time and (b) kinetic fit for the degradation of RhB.
Catalysts 07 00243 g005
Figure 6. UV–vis absorption spectra of RhB in (a) TiOF2 and (b) NaOH-modified TiOF2 suspension under visible light.
Figure 6. UV–vis absorption spectra of RhB in (a) TiOF2 and (b) NaOH-modified TiOF2 suspension under visible light.
Catalysts 07 00243 g006

Share and Cite

MDPI and ACS Style

Hou, C.; Liu, W.; Zhu, J. Synthesis of NaOH-Modified TiOF2 and Its Enhanced Visible Light Photocatalytic Performance on RhB. Catalysts 2017, 7, 243. https://doi.org/10.3390/catal7080243

AMA Style

Hou C, Liu W, Zhu J. Synthesis of NaOH-Modified TiOF2 and Its Enhanced Visible Light Photocatalytic Performance on RhB. Catalysts. 2017; 7(8):243. https://doi.org/10.3390/catal7080243

Chicago/Turabian Style

Hou, Chentao, Wenli Liu, and Jiaming Zhu. 2017. "Synthesis of NaOH-Modified TiOF2 and Its Enhanced Visible Light Photocatalytic Performance on RhB" Catalysts 7, no. 8: 243. https://doi.org/10.3390/catal7080243

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