Anda di halaman 1dari 5

Chemical Engineering Science 63 (2008) 5066 5070 www.elsevier.

com/locate/ces

Hydrothermal synthesis of TiO2 photocatalysts in the presence of NH4 F and their application for degradation of organic compounds
Kohsuke Mori, Keiichi Maki, Shinichi Kawasaki, Shuai Yuan, Hiromi Yamashita
Division of Materials and Manufacturing Science, Graduate School of Engineering, Osaka University, 2-1 Yamada-oka, Suita, Osaka 565-0871, Japan Received 13 April 2007; received in revised form 22 June 2007; accepted 26 June 2007 Available online 30 June 2007

Abstract TiO2 photocatalysts were synthesized by hydrothermal method from tetraisopropyl orthotitanate (TPOT) in the presence of NH4 F with different NH4 F/Ti molar ratios (HT-TiO2 (F0, 0.01, 0.05, 0.25, and 1)). The formation of well-crystallized anatase phase of TiO2 and suppression of phase transition to rutile were observed even at high calcination temperature. The amount of employed NH4 F inuenced on the crystal size and surface area. XPS analysis showed that F ions coordinated with Ti in the lattice. In comparison to the commercial TiO2 powder (P-25), the HT-TiO2 samples with high F ion contents exhibited high absorption in the UVvisible range with a shift to the longer wavelength in the band gap transition. The measurement of the water adsorption ability suggested that the F ion doping did not signicantly inuence on their hydrophobicity. The HT-TiO2 samples exhibited high photocatalytic activity for the degradation of i-BuOH diluted in water. The photocatalytic activities were apparently affected by surface area and crystallinity of TiO2 phase depending on the amount of employed NH4 F. 2007 Elsevier Ltd. All rights reserved.
Keywords: Photocatalyst; TiO2 ; NH4 F; Hydrothermal synthesis; Photocatalytic degradation

1. Introduction Photocatalytic decomposition of organic pollutants on semiconductor surfaces is of vital interest with respect to both fundamental understanding and potential practical utilization (Yamashita and Anpo, 2003; Horikoshi et al., 2002, Minero et al., 2000a,b). Titanium dioxide TiO2 is a promising photocatalyst exploiting low cost, nontoxity as well as strong oxidation ability under UV light irradiation (Mills and Hunte, 1997; Fujishima et al., 2002). Especially, TiO2 thin lm can be used in a wide variety of practical applications such as selfcleaning materials in automobiles, buildings, and household glazing (Wang et al., 1997). In spite of the above characteristics, most of applications of TiO2 photocatalyst so far are limited to UV-light irradiation because the light absorption edge of pure TiO2 is less than 380 nm. Moreover, hydrophilic nature of TiO2 surface prevents the efcient transfer of hydrophobic

Corresponding author. Tel./fax: +81 6 6879 7457.

E-mail address: yamashita@mat.eng.osaka-u.ac.jp (H. Yamashita). 0009-2509/$ - see front matter 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.ces.2007.06.030

organic compounds from solution or atmosphere to TiO2 particle surface, which resulted in the decrease of photocatalytic activity. Therefore, the great demand for the development of TiO2 photocatalysts with enhanced activity is still high. During the past decade, TiO2 photocatalysts operating effectively under visible-light irradiation have been presented by the implantation of transition metal cations, e.g. Fe, Cr, and V (Yamashita and Anpo, 2004). Another noble metals loading, such as Pt and Ag, can decrease the recombination of photogenerated electrons and holes, which promote interfacial electron transfer (Zou et al., 2004). Asahi et al. (2001) also demonstrated that the doping of nitrogen into TiO2 provided the visible-light sensitivity. The induced visible-light sensitivity of the N-doped TiO2 was due to the narrowing of the band gap by hybridization of the N 2p and O 2p orbitals, or formation of an isolated narrow N 2p band above the valence band. As an alternative method, the doping of F ion within the TiO2 lattice has received much attention. TiO2 doped with F in the lattice using NaF as a uorine source showed enhanced photocatalytic activity (Park and Choi, 2004; Minero et al., 2000a,b). F -doped TiO2 were also prepared by spray pyrolysis method (Li et al., 2005). It has been reported that the improved photocatalytic

K. Mori et al. / Chemical Engineering Science 63 (2008) 5066 5070

5067

activity could be ascribed to the increase in the anatase crystallinity induced by F ion. The introduction of F ion is also promising to reduce the hydrophilic nature of TiO2 particle surface, which could enhance the afnity of TiO2 surface toward organic compounds by the replacement of O.H group by F (Yuan et al., 2001). The modication of hydrophilic TiO2 surface also could be achieved by HF or F2 treatment (Pong et al., 1995). In this study, we designed TiO2 photocatalyst in the presence of NH4 F under hydrothermal conditions (HT-TiO2 ). The effects of initial NH4 F contents on the characteristics of TiO2 were investigated by several physicochemical characterizations. Moreover, its successful utilization was investigated for the photocatalytic degradation of isobutanol (i-BuOH) diluted in water. 2. Experimental Tetraisopropyl orthotitanate (TPOT: Ti(OC3 H7 )4 ), ethanol, NH4 F, and i-BuOH are commercial products. All chemicals were used as received. The synthesis of HT-TiO2 were performed by hydrothermal method using TPOT as the starting material in a mixed NH4 F.H2 Oethanol solution. The molar ratio of the mixture was as follows: Ti : H2 O : EtOH = 1 : 5 : 5. The above mixture was transferred into an autoclave and kept at 433 K for 48 h. After centrifugation of the resultant mixture, followed by washing, drying, and calcination at 823 K for 5 h, HT-TiO2 was obtained. NH4 F/Ti molar ratios were set at 0, 0.01, 0.05, 0.25, and 1, respectively. The products were denoted as HT-TiO2 (F0), HT-TiO2 (F0.01), HT-TiO2 (F0.05), HT-TiO2 (F0.25), and HT-TiO2 (F1), respectively. As a reference sample, TiO2 was also prepared by the conventional solgel method. The mixture of TPOT, H2 O, and EtOH (1 : 5 : 5) was stirred at room temperature for 24 h, and then heated at 358 K for 24 h. After centrifugation of the resultant mixture, followed by drying, and calcination at 823 K for 5 h, solgel TiO2 was obtained. X-ray diffraction patterns of all samples were measured by Rigaku RINT2500 diffractometer with Cu K radiation ( = By using an ASAP 2000 system (Shimadzu), the 1.5406 A). BET method was applied for the determination of the specic surface area. The diffuse reectance absorption spectra were recorded with Shimadzu UV-2200A photospectrometer. X-ray photoelectron spectroscopy was recorded with JEOL microprobe system (JPS-9010) using the Mg K line. SEM micrograph was obtained with Hitachi S-5200. The photocatalytic activity of samples was evaluated by the photodegradation of i-BuOH. A 0.05 g of catalyst was dispersed in 25 ml of aqueous i-BuOH solution (2.61 mmol l1 ). After stirring under dark conditions for 30 min, the solution was bubbled by oxygen for another 30 min. Then the solution was irradiated using UV light ( > 280 nm) from a 100 W high-pressure Hg lamp. The progress of the reactions was monitored by gas chromatography (GC-14B, Shimadzu). The photocatalytic activity was estimated from the initial decrease in the concentration of i-BuOH after preadsorption of i-BuOH on the catalyst under dark conditions. Water adsorption isotherms of the

catalysts were measured at 293 K using a conventional vacuum system. 3. Results and discussion Fig. 1 shows the XRD patterns of hydrothermally synthesized and calcined at 823 K HT-TiO2 samples with different NH4 F loading levels, together with the reference TiO2 prepared by the conventional solgel method. In all the HT-TiO2 samples, very sharp peaks corresponding to the anatase phase of TiO2 could be observed. No other phases such as rutile and brookite could be detected. The crystallinity of the anatase phase (I /Isol.gel ) in HT-TiO2 was estimated by the intensity of (1 0 1) reection relative to that of the solgel TiO2 and the results are given in Table 1. The HT-TiO2 (0.25) showed the highest I /Isol.gel value. By applying the Scherrer equation for (1 0 1) reection, the average crystallite size of samples can be determined to be 20 nm (HT-TiO2 (F0)), 18 nm (HT-TiO2 (F0.01)), 20 nm (HT-TiO2 (F0.05)), 24 nm (HT-TiO2 (F0.25)), and 21 nm (HT-TiO2 (F1)), respectively. The crystallite sizes were not dependent on the amount of NH4 F employed, but the HT-TiO2 (0.01) has the smallest anatase crystallite size among the samples prepared. The results of the BET surface areas reected the crystallite sizes determined by XRD, as summarized in Table 1. The surface area decreased with increasing the crystal size. On the contrary, additional peaks corresponding to the rutile phase at 27.5 and 36.0 were observed in the solgel TiO2 (Fig. 1a). It should be noted that the present hydrothermal synthesis using TPOT as a Ti source in the presence of NH4 F inhibits the phase transformation from anatase to rutile. In order to further verify the thermal stability of the HT-TiO2 , the samples were calcined at a higher temperature of 923 K and the XRD patterns of them are given in Fig. 2. The result of the TiO2 prepared by the solgel method was also depicted as

(f ) HT-TiO2(1)

(e) HT-TiO2(0.25)

Intensity / a.u.

(d) HT-TiO2(0.05)

(c) HT-TiO2(0.01)

(b) HT-TiO2(0)

(a) sol-gel TiO2

20

30

40 2 / degree

50

60

Fig. 1. XRD patterns of TiO2 prepared by the solgel method and hydrothermally synthesized HT-TiO2 samples with different amount of NH4 F after calcination at 823 K.

5068

K. Mori et al. / Chemical Engineering Science 63 (2008) 5066 5070

Table 1 Crystal size, surface area, and amount of H2 O adsorption of the HT-TiO2 samples calcined at 823 K

HT-TiO2 HT-TiO2 HT-TiO2 HT-TiO2 HT-TiO2


a The b The

(0) (0.01) (0.05) (0.25) (1)

1.03 1.05 1.20 1.44 1.17

55 59 48 33 47

18 17 20 16 15

intensity of (1 0 1) reection relative to that of the solgel TiO2 . values at P0 /P1 = 1.

Absorbance / a.u. 695

Sample

I /Isol.gel a

Surface area (m2 g1 )

H2 O adsorptionb ( mol m2 )

690 685 Binding Energy / eV

680

(f) HT-TiO2(1)

Fig. 3. XPS spectrum of F1S in HT-TiO2 (1).

(e) HT-TiO2(0.25)

Intensity / a.u.

(d) HT-TiO2(0.05)

(c) HT-TiO2(0.01)

(b) HT-TiO2(0)

(a) sol-gel TiO2

20

30

40 2 / degree

50

60

Fig. 2. XRD patterns of TiO2 prepared by the solgel method and hydrothermally synthesized HT-TiO2 samples with different amount of NH4 F after calcination at 923 K.

reference sample. The HT-TiO2 (F0) and HT-TiO2 (F0.01) showed the formation of the mixture of anatase and rutile phases. With increasing the NH4 F contents, the phase transfer to rutile was completely suppressed and all peaks could be indexed as the anatase phase of crystalline TiO2 . On the other hand, the rutile phase was dominant in the solgel TiO2 with a small amount of anatase phase. Similar phenomenon was also reported in the F-doped TiO2 powder prepared by hydrolysis of titanium tetraisopropoxide in a mixed NH4 F.H2 O solution (Li et al., 2005). These results clearly indicate that F ions enhance the crystallization of anatase phase, and meanwhile suppress the crystallization into rutile phase. Fig. 3 shows the F1S XPS spectrum of the HT-TiO2 (F1). Generally, the F1S binding energy at around 684 eV corresponds to the adsorbed F ions on TiO2 surface, while that at 688 eV is ascribed to the F ions in the TiO2 lattice (Wang and Sherwood, 2004). Although the intensity of the peak of F1S was weak in the HT-TiO2 (F1), one major peak ascribed to the F ions in the TiO2 lattice exhibited strong evidence of F ions was in-

corporated into TiO2 crystals. On the other hand, no noticeable signals assignable to N1S binding energy could be observed in the HT-TiO2 (F1), suggesting that N atoms originating from the NH4 F scarcely present. The differences in the H2 O adsorption capability were compared and the results are listed in Table 1. The F ion doping did not signicantly affect the amount of adsorbed H2 O molecules, which is ascribed to that most of the F ions exist in the TiO2 lattice rather than simply being adsorbed on the surface, as conrmed by XPS analysis. However, the amount of adsorbed H2 O molecules slightly decreased on the HT-TiO2 (F0.25) and HT-TiO2 (F1), which may indicate that the hydrophobic properties were enhanced by the replacement of O.H group by F . Fig. 4 shows the SEM image of the HT-TiO2 (F0) and HT-TiO2 (F0.25) samples calcined at 823 K. The gross morphology of HT-TiO2 remained unchanged by the addition of the NH4 F. In spite of the smaller surface area of the HT-TiO2 (F0.25) sample compared to the HT-TiO2 (F0), the noticeable agglomeration of TiO2 particles could not be observed. F ions doping clearly inuenced on the light absorption characteristics. UVvis spectra of the HT-TiO2 samples are shown in Fig. 5. Compared with the pure TiO2 (P-25), the HT-TiO2 (F0.25) and HT-TiO2 (F1) samples showed a stronger absorption in the UV range and a slight shift to visible-light range (> 430 nm). HT-TiO2 (F0) and HT-TiO2 (F0.01) samples exhibited similar absorption edge to that of the pure TiO2 . It has been reported that the F ion doping does not cause a shift in the fundamental absorption edge of TiO2 (Ho et al., 2006). This results is consistent with the theoretical band calculations for F ions doped TiO2 (Morikawa et al., 2001; Yamaki et al., 2003). Although we could not detect the nitrogen atoms by the XPS analysis, an extremely small amount of N contamination from NH4 F in the HT-TiO2 (F0.25) and HT-TiO2 (F1) samples might have shifted the band edges to the longer wavelength. Increase of the absorption intensity as well as response range would increase the number of photogenerated electrons and holes to participate in the photocatalytic reaction. The reaction rates for the photocatalytic degradation of i-BuOH diluted in water were shown in Fig. 6. The reaction

K. Mori et al. / Chemical Engineering Science 63 (2008) 5066 5070

5069

0.08

f c

d
Kubelka-Munk

10

-3

0.06 Kubelka-Munk

1 a b c d e

b 0.04 a

0 400

0.02

450 500 Wavelength / nm

550

0 400 Wavelength / nm
Fig. 5. UVvis spectra of (a) TiO2 (P-25), (b) HT-TiO2 (0), (c) HT-TiO2 (0.01), (d) HT-TiO2 (0.05), (e) HT-TiO2 (0.25), and (f) HT-TiO2 (1).

500

P25 HT-TiO2(1) HT-TiO2(0.25) HT-TiO2(0.05) HT-TiO2(0.01) HT-TiO2(0) 0


Fig. 4. SEM image of (a) HT-TiO2 (0), and (b) HT-TiO2 (0.25) calcined at 823 K.

10

15
1

20

Conversion of i-BuOH / molh

Fig. 6. Photocatalytic degradation of i-BuOH diluted in water on P25 and various HT-TiO2 synthesized with different amount of NH4 F.

rate of the commercially available P25 was determined to be 15 mol h1 . The HT-TiO2 (F0.05) was the most active photocatalyst among the catalyst tested, and the HT-TiO2 (F0.25) showed similar activity to that of the HT-TiO2 (F0.01). HT-TiO2 (F0) and HT-TiO2 (F1) apparently exhibited decreased photocatalytic activities. The observed activities of the photocatalysts are combined effects of many factors including phase structure, surface area, crystallinity. The relatively high photocatalytic activity of the HT-TiO2 (F0.01) is partially attributed to its large surface area and small crystallite size. In the case of the HT-TiO2 (F0.25), high crystallinity of the anatase phase is a main factor for the high activity (Yu et al., 2002; Murakami et al., 2007). A photoabsorption ability in the wider range observed in the UVvis spectrum may also plays an important role. The highest activity of the HT-TiO2 (F0.05) is attributed to both relatively high crystallinity as well as high surface area. It has been reported that F doping in the lattice could convert some Ti4+ to Ti3+ originated from the charge compensation caused by the replacement of F ions with O2 ions in the lattice (Yu et al., 2002). The Ti3+ could trap photogenerated holes, which resulted in charge recombination in

bulk and lower photocatalytic activity. It can be said that an excessive F doping in the HT-TiO2 (F1) plays a negative role in photocatalysis. 4. Conclusions F -doped TiO2 photocatalysts were prepared by a hydrothermal method in the presence of NH4 F. This is a promising technique to synthesize the well-crystallized anatase TiO2 , suppressing the transition to rutile phase. The amount of employed NH4 F inuenced on the crystal size and surface area as well as photocatalytic activity. The HT-TiO2 (F0.05) was demonstrated to be most active for the degradation of i-BuOH diluted in water. The photocatalytic activity of the HT-TiO2 samples is dominated by the crystallinity of anatase phase, and surface area. Acknowledgements The present work is supported by the Grant-in-Aid for Scientic Research (KAKENHI) from Ministry of Education, Culture, Sports, Science and Technology (No. 1734036),

5070

K. Mori et al. / Chemical Engineering Science 63 (2008) 5066 5070 Murakami, S.-i., Kominami, H., Kera, Y., Ikeda, S., Noguchi, H., Uosaki, K., Ohtani, B., 2007. Evaluation of electronhole recombination properties of titanium(IV) oxide particles with high photocatalytic activity. Research on Chemical Intermediates 33, 285296. Park, J.S., Choi, W., 2004. Enhanced remote photocatalytic oxidation on surface-uorinated TiO2 . Langmuir 20, 1152311527. Pong, T.K., Besida, J., ODonnell, T.A., Wood, D.G., 1995. A novel uoride process for producing TiO2 from titaniferous ore. Industrial and Engineering Chemistry Research 34, 308313. Wang, Y.Q., Sherwood, P.M.A., 2004. Studies of carbon nanotubes and uorinated nanotubes by X-ray and ultraviolet photoelectron spectroscopy. Chemistry of Materials 16, 54275436. Wang, R., Hashimoto, K., Fujishima, A., Chikuni, M., Kojima, E., Kitamura, A., Shimohigoshi, M., Watanabe, T., 1997. Light-induced amphiphilic surfaces. Nature 388, 431432. Yamaki, T., Umebayashi, T., Sumita, T., Yamamoto, S., Maekawa, M., Kawasuso, A., Itoh, H., 2003. Fluorine-doping in titanium dioxide by ion implantation technique. Nuclear Instrument and Methods in Physics Research Section B 206, 254258. Yamashita, H., Anpo, M., 2003. Local structures and photocatalytic reactivities of the titanium oxide and chromium oxide species incorporated within micro- and mesoporous zeolite materials: XAFS and photoluminescence studies. Current Opinion in Solid State and Materials Science 7, 471481. Yamashita, H., Anpo, M., 2004. Application of an ion beam technique for the design of visible light-sensitive, highly effcient and highly selective photocatalysts: ion-implantation and ionized cluster beam methods. Catalysis Survey from Asia 8, 3545. Yu, J.C., Yu, J., Ho, W., Jiang, Z., Zhang, L., 2002. Effects of F doping on the photocatalytic activity and microstructures of nanocrystalline TiO2 powders. Chemistry of Materials 14, 38083816. Yuan, Q., Ravikrishna, R., Valsaraj, K.T., 2001. Reusable adsorbents for dilute solution separation. 5. Photodegradation of organic compounds on surfactant-modied titania. Separation and Purication Technology 24, 309318. Zou, J.J., Liu, C.J., Yu, K.L., Cheng, D.G., Zhang, Y.P., He, F., Du, H.Y., Cui, L., 2004. Highly efcient Pt/TiO2 photocatalyst prepared by plasmaenhanced impregnation method. Chemical Physics Letters 400, 520523.

(No. 17360388) and (No. 18656238). This work is partly performed under the project of collaborative research at the Joining and Welding Research Institute (JWRI) of Osaka University. References
Asahi, R., Morikawa, T., Ohwaki, T., Aoki, K., Taga, Y., 2001. Visiblelight photocatalysis in nitrogen-doped titanium oxides. Science 293, 269273. Fujishima, A. Rao, T.N., Tryk, D.A., 2000. Titanium dioxide photocatalysis. Journal of Photochemistry and Photobiology C: Photochemistry Reviews 1, 121. Ho, W., Yu, C.Y., Lee, S., 2006. Synthesis of hierarchical nanoporous F-doped TiO2 spheres with visible light photocatalytic activity. Chemical Communications 11151117. Horikoshi, S., Hidaka, H., Serpone, N., 2002. Environmental remediation by an integrated microwave/UV-illumination method. 1. Microwaveassisted degradation of rhodamine-B dye in aqueous TiO2 dispersions. Environmental Science and Technology 36, 13571366. Li, D., Haneda, H., Hishita, S., Ohashi, N., Labhsetwar, N.K., 2005. Fluorinedoped TiO2 powders prepared by spray pyrolysis and their improved photocatalytic activity for decomposition of gas-phase acetaldehyde. Journal of Fluorine Chemistry 126, 6977. Mills, A., Hunte, S.L., 1997. An overview of semiconductor photocatalysis. Journal of Photochemistry and Photobiology A: Chemistry 108, 135. Minero, C., Mariella, G., Maurino, V., Pelizzetti, E., 2000a. Photocatalytic transformation of organic compounds in the presence of inorganic anions. 1. Hydroxyl-mediated and direct electron-transfer reactions of phenol on a titanium dioxide-uoride system. Langmuir 16, 26322641. Minero, C., Mariella, G., Maurino, V., Pelizzetti, E., 2000b. Photocatalytic transformation of organic compounds in the presence of inorganic ions. 2. Competitive reactions of phenol and alcohols on a titanium dioxideuoride system. Langmuir 16, 89648972. Morikawa, T., Asahi, R., Ohwaki, T., Aoki, K., Taga, Y., 2001. Band-gap narrowing of titanium dioxide by nitrogen doping. Japanese Journal of Applied Physics 40, 561563.

Anda mungkin juga menyukai