Sie sind auf Seite 1von 9

Materials Research Bulletin 43 (2008) 21872195 www.elsevier.

com/locate/matresbu

Synthesis and photocatalytic properties of TiO2 nanostructures


X.H. Xia a,*, Y. Liang a, Z. Wang b, J. Fan a, Y.S. Luo a, Z.J. Jia a,*
a

Institute of Nano-science and Technology, Huazhong Normal University, Wuhan 430079, China b Department of Physics Science and Technology, Wuhan University, Wuhan 430072, China Received 5 June 2006; received in revised form 11 August 2007; accepted 27 August 2007 Available online 31 August 2007

Abstract TiO2 particles, rods, owers and sheets were prepared by hydrothermal method via adjusting the temperature, the pressure and the concentration of TiCl4. The as-prepared TiO2 powders were characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), X-ray photoelectron spectroscopy (XPS), UVvis diffuse reectance spectra and N2 adsorptiondesorption measurements. It was found that pressure is the most important factor inuencing the morphology of TiO2. The photocatalytic activity of the products was evaluated by the photodegradation of aqueous brilliant red X-3B solution under UV light. Among the as-prepared nanostructures, the ower-like TiO2 exhibited the highest photocatalytic activity. # 2007 Elsevier Ltd. All rights reserved.
Keywords: A. Nanostructures; A. Semiconductors; B. Chemical synthesis; C. Electron microscopy; C. X-ray diffraction; D. Catalytic properties

1. Introduction In recent years, there is increasing interest in the development of semiconductor oxide photocatalysts for environmental protection [1]. Among the various semiconductor oxide photocatalysts, TiO2 is widely used because of its strong oxidizing power, non-toxicity and long-term stability [2]. It is important to control the morphology and the crystal structure of TiO2 since its photocatalytic activity is strongly dependent on the crystal structure, the crystallite size and the specic surface area [3,4]. Various techniques such as the solgel method [5], the chemical vapor decomposition method [6,7], the hydrothermal technique [8], the reversed micelle method [9] were used to produce TiO2 nanostructures. The hydrothermal reactions are carried out in a closed system and the products can be well controlled by the temperature, the pressure and the concentration of the chemical species [1013], which results in the popularity of the hydrothermal method in the synthesis of TiO2. Particle and rod-like TiO2 structures are easy to be synthesized by this method, while the ower-like TiO2 prepared by this method is rarely reported in literature. And a systematic study on the synthesis of TiO2 by hydrothermal method is lacking. The purpose of this paper is to report a systematic study on the controlled synthesis of TiO2 by hydrothermal method, as well as to investigate the photocatalytic activity of the as-prepared TiO2 in photodegradation of aqueous brilliant red X-3B solution. The results are important for further understanding of the morphology control and the crystal growth of TiO2 and other semiconductor materials.

* Corresponding authors. Tel.: +86 27 67 861185; fax: +86 27 67 861185. E-mail address: xiaxh@phy.ccnu.edu.cn (X.H. Xia). 0025-5408/$ see front matter # 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.materresbull.2007.08.026

2188

X.H. Xia et al. / Materials Research Bulletin 43 (2008) 21872195

2. Experimental All the chemicals were of analytical grade. Titanium tetrachloride (98% TiCl4, Tianjin Fuchen Chemical Reagent factory, China) and hydrochloric acid (38% HCl, Kaifeng Dongda Chemical Ltd. Co., China) were used as the starting materials without any additives. In a typical procedure, 1 mL of TiCl4 was added dropwise into 10 mL of hydrochloric acid at room temperature. The obtained transparent solution was then diluted by adding distilled water under vigorous stirring. The concentration of the TiCl4 was controlled by adjusting the amount of water. The pH values of the solutions were measured by an acidimeter (pHS-2, Shanghai Rex Instruments Factory) at about one. After the dilution, 250 mL of the solution was placed into an autoclave and the initial pressure of the hydrothermal system was formed by introducing air into the autoclave. The hydrothermal reactions were then maintained for 5 h at xed temperatures. Nine samples AI were prepared as above and the detailed synthesis conditions of them are shown in Table 1. The nal white precipitations were taken out after the system cooled down to room temperature and washed with deionized water by centrifugation till the pH value of water was below 7.5. The as-prepared samples were then dried in air at 80 8C overnight. The morphologies of the products were investigated by scanning electron microscopy (SEM, JEOL-6700F) and transmission electron microscopy (TEM, JEM-2010). Powder X-ray diffraction (XRD) patterns were obtained by a Y200 diffractometer (D/max 30 kV) using Cu Ka radiation (l = 0.154178 nm). The N2-adsorption BrunauerEmmett Teller (BET) surface areas were measured by N2-adsorption at 77 K (Tristar3000, Micromeritics) and the X-ray photoelectron spectroscopy (XPS) measurements were done with a Kratos XSAM800 XPS system with Cu Ka source and a charge neutralizer, all the binding energies were referenced to the C 1s peak at 284.4 eV of the surface adventitious carbon. The photocatalytic activities of the as-prepared samples were tested in the photodegradation of aqueous brilliant red X-3B solution after sintered at 450 8C for 2 h. The reactive brilliant red X-3B (98%) (Jining Dye Manufacture of Shengdong, China) was used directly without further purication. Molecule structure of the X-3B dye was shown in Fig. 1. In a typical procedure, the investigated photocatalyst sample was added into a beaker in which 500 mL of the brilliant red X-3B solution (100 mg/L) was added previously. The concentration of the photocatalyst was xed at 400 mg/L. The pH value of the reaction medium was about seven. Prior to the photo-oxidation, the suspension with X3B and the photocatalyst was magnetically stirred in the dark condition for 1 h to establish an adsorptiondesorption equilibrium in an airproof condition. The suspension was then transferred into the inner tube of the photoreactor and irradiated with a 375 W high-pressure mercury lamp (l = 365 nm). Air was bubbled into the reaction suspension at a ow rate of 120 mL/min throughout the experiments. Samples were obtained at an interval of 5 min and the absorbencies of the samples were checked on lambda UVvis spectrophotometer (UV-2500, Shimadzu, Japan) after centrifugation. 3. Results and discussion Fig. 2 shows the XRD patterns of samples AI. For samples AF, they are anatase TiO2 (JCPDS no. 21-1272) and for samples GI, they are rutile TiO2 (JCPDS no. 77-0441). No characteristic peaks of impurities such as brookite TiO2
Table 1 Reaction conditions, surface areas, adsorption rates and photocatalytic degradation rates of sample AI Sample A B C D E F G H I T (8C) 80 80 80 160 160 160 320 320 320 P (MPa) 0 0.8 0.8 0 0.8 0.8 0 0.8 0.8 C (M) 0.5 0.5 1 0.5 0.5 1 0.5 0.5 1 SBET (m2/g) 43.09 40.50 29.17 41.69 63.35 31.96 19.17 17.32 10.26 A (%) 12.37 11.59 2.93 10.76 15.27 9.57 9.22 4.16 2.05 D (%) 96.7 66.76 54.9 79.84 98.94 63 41.01 38.7 22.9

T, P, C, SBET, A and D represents temperature, the initial pressure, the concentration of TiCl4, BET surface area, adsorption rate and the degradation rate of X-3B, respectively.

X.H. Xia et al. / Materials Research Bulletin 43 (2008) 21872195

2189

Fig. 1. Chemical formula of X-3B.

Fig. 2. The XRD patterns of samples AI.

and Ti(OH)4 are detected. The widths of the diffraction peaks of samples AI are in the following orders: A > B > C; D > E > F; and G > H > I and the intensities of the diffraction peaks follow exactly the contrary orders. As we can see from Table 1, the pressure and the concentration of TiCl4 is also in the same orders. It is well accepted that the broadening of the diffraction peaks reects the decrease of the particle size and the intensity of the diffraction peaks reects the crystallization of the samples [14]. The results suggest that the particle size and the crystallization of TiO2 both increase with the increase of pressure and the concentration of TiCl4. Fig. 3 shows the electron micrograph images of samples AI. It can be seen from Fig. 3(a) that in sample A there are small particles about 30 nm in diameter, and the particles badly aggregated together. Rod-like structures were obtained both in samples B and C. The rods in sample B are about 60 nm in diameter and 200 nm in length as shown in Fig. 3(b). While the diameters of the rods in sample C are much bigger which are about 200 nm and the lengths are about 500 nm as shown in Fig. 3(c). The SEM images of samples DF are displayed in Fig. 3(d), (e) and (g), respectively, they are all ower-like structures with TiO2 nanorods radically growing from the center. In Fig. 3(d), small ower-like structures as well as particles were obtained. The short nanorods are about 10 nm in diameter and 30 nm in length radiating out. The particles are about 100 nm in size. Prolonged reaction time could not facilitate the growth of the particles into owers, indicating that the pressure is a more crucial factor for the formation of owers rather than the reaction time. The rods radiated from the center in sample E are about 30 nm in diameter and 100200 nm in length as shown in Fig. 3(e). Fig. 3(f) is the TEM image of sample E, which demonstrates the ower-like structure. The rods radiating from the center of the owers in sample F are square columns rather than spine rods. The columns have a diameter of about 80 nm and a length of about 300 nm. The TEM pictures of samples GI are shown in Fig. 3(h)(j). The products in sample G are very thin sheets about 80 nm in size. The products in sample H are also thin nanosheets yet a little bigger, which are about 100 nm in size as indicated in Fig. 3(i). The nanosheets are about 200 nm in sample I and they are thicker than those in samples G and H. From the results we can deduce that the additional pressure in the hydrothermal system at 80 8C and 160 8C is helpful for the formation of TiO2 nanorods and owers, respectively, while the initial pressure in the 320 8C

2190

X.H. Xia et al. / Materials Research Bulletin 43 (2008) 21872195

Fig. 3. SEM images of sample A (a), sample B (b), sample C (c), sample D (d), sample E (e), sample F (g) and TEM images of sample E (f), sample G (h), sample H (i) and sample I (j).

hydrothermal system did not affect the morphology of TiO2 so much. This can be interpreted that the pressure formed in the 320 8C system by itself is high enough for the formation of the nanosheets. It can be seen from Table 1 that the Ti4+ concentrations are in the following order: B < C; E < F; and H < I, and from Fig. 2 we observed that the particle size are in the same order. Then we can deduce that the increase of the Ti4+ concentration results in the increase of the crystal size. It is well known that TiCl4 is very easy to hydrolysis in water at room temperature. The hydrolysis of TiCl4 procedures in the following steps as (1) and (2) [15]: TiCl4 H2 O ! TiOH3 H 4Cl TiOH3 ! TiO2 H TiO2 H2 O ! TiO2 2H The rst step was usually very fast and the additive HCl in our experiment can suppress the hydrolysis rate, which allows TiCl4 to hydrolyze at an appropriate rate. In our experiment, the more the addition of the hydrochloric acid, at (1) (2)

X.H. Xia et al. / Materials Research Bulletin 43 (2008) 21872195

2191

higher temperature the TiCl4 began to hydrolysis and the product disperses better. This is because that the hydrolysis of TiCl4 is inhibited due to the acidity of the aqueous solution. The more the hydrochloric acid was added, the more H+ in the solution and the slower the hydrolysis reaction proceeds, thus the higher temperature was required to boost the hydrolysis reaction. When TiCl4 reacts with water, the Ti4+ increases its co-ordination to six by using its vacant d-orbits to accept oxygen lone pairs from nucleopholic ligands. These sixfold structural units undergo condensation and become the octahedral that are incorporated into the nal precipitate structure [16,17]. Both anatase and rutile phases belong to the tetragonal crystal system, consisting of TiO6 octahedra as a basic structural unit. Their crystalline structures differ in the assembly of the octahedra chains. Arrangement of octahedra through face sharing initiates the anatase while the edge sharing leads to the rutile phase. In the initial stage of experiment, a three-dimensional cluster would be formed, which could act as the nucleus of TiO2 crystals and further develop into crystallite. The further development will be controlled by the outer chemical and physical conditions. At a lower temperature of 80 8C, TiCl4 hydrolyzed slowly and the TiO2 nucleus can grow slowly and freely in all the directions, then the particle-like structures could form freely. With the addition of 0.8 MPa pressure, the TiO6 octahedras are prefer to grow along the axis parallel to [0 1 0] direction, along which the TiO octahedras share one face, forming the anatase crystals. When the reaction temperature increased to 160 8C, the reaction rates of steps (1) and (2) became faster. It promoted large generation of TiO2 nucleus in a short time resulting in the growth of owers with rods radiated from the center. Higher pressure at 160 8C results in bigger owers with bigger rods radiating from the center. However, the growth along the [0 1 0] direction at an excessive high temperature as 320 8C was restrained and the TiO6 octahedras share a pair of opposite edges in the [0 0 1] direction, forming the sheet-like rutile TiO2. With the increase of TiCl4 concentration, the particle size and the crystallization of the products increased. This is because that both steps (1) and (2) go along faster at bigger concentration of TiCl4. Fig. 4 is the schematic illustration of the growth mechanisms for TiO2 nanocrystals prepared at different conditions. The TiO2 crystals could present spontaneously and radially preferential growth from the active sites on the nucleus due to the anisotropic growth habit of TiO2 and the outer chemical and physical conditions including temperature, pressure and concentration of the reaction solution [18]. X-3B is a common chemical used extensively in a variety of industrial application. Therefore, we chose it as a model contaminate chemical in our photocatalytic reaction. In order to conrm the photocatalysis of the TiO2 nanostructures, the direct photolysis of X-3B was also investigated under the same reaction condition without photocatalyst. Fig. 5 shows the UVvis absorption spectra of X-3B before and after direct photolysis. Only 2.17% of X-3B disappeared after 30 min photolysis. It means that the direct photolysis was a weak working factor to affect the photodegradation of X-3B dye, and the photocatalysis by the TiO2 nanostructures was the main reason to cause the X3B degradation.

Fig. 4. Schematic illustration of the growth mechanisms for TiO2 nanocrystals.

2192

X.H. Xia et al. / Materials Research Bulletin 43 (2008) 21872195

Fig. 5. UVvis absorption spectra of X-3B before and after direct photolysis.

It is signicant to note that the adsorption of X-3B on the surface of TiO2 was very important, only after X-3B was rst adsorbed on the surface of the catalysts, could the X-3B be oxidized by the radical species [19]. The adsorption rates and the surface areas of samples AI are shown in Table 1. From the table we could see that the bigger surface area of the sample is, the higher adsorption rate is. The adsorption of sample E is the highest, and the adsorptions of the

Fig. 6. UVvis absorption spectra of X-3B after 30 min photcatalysis with samples A and E as the photocatalysts.

X.H. Xia et al. / Materials Research Bulletin 43 (2008) 21872195

2193

samples prepared at 320 8C are lower than that prepared at 80 8C and 160 8C. The degradation rates of aqueous brilliant red X-3B on samples AI are also displayed in Table 1. Among the as-prepared TiO2 samples, the degradation rate of sample E is the highest and sample I is the lowest. The degradation rates of the samples prepared at 80 8C and 320 8C decreased with the increase of pressure and the concentration of TiCl4. While at 160 8C, sample E prepared with 0.8 MPa pressure exhibits higher degradation rate both than sample F prepared at bigger concentration of TiCl4 and sample D prepared at smaller pressure, this may due to the high surface area of sample E. Fig. 6 shows the UVvis absorption spectra of X-3B after 30 min photcatalysis with samples A and E as the photocatalysts. As to Reactive Brilliant Red X-3B, within visible region (l = 537 nm), its absorption band is relevant to the whole conjugated structure. Just the pp* transition of electrons in the azo-group connecting phenyl and naphthyl causes the band. Within near ultraviolet region (l = 217, 298 nm), its absorption band results from the unsaturated system of benzene and naphthalene ring. In comparison with the UVvis characteristic absorption peaks of X-3B molecule, the apparent decrease of the absorption intensities indicated the photocatalytic capability of the photocatalysts to degrade X-3B. The decolorization resulted from the destruction of the conjugated azo double bonds and the other un-saturated system of the dyestuff [20]. The characteristic absorption peaks of X-3B at 298 nm and 537 nm decreased stepwise in both the two graphs. In the graph of sample A, the two peaks still appeared after 30 min photocatalysis though it was much lower. While in sample E, the two peaks almost disappeared after 30 min photocatalysis. It indicates that the azoN = Nconjugate structure and the un-saturated benzene and naphthalene rings of X-3B were almost completely destroyed by sample E. And the photocatalytic efciency of sample E is higher than that of sample A. Fig. 7(a) shows the XPS spectra of the TiO2 owers, particles, rods and sheets after calcined at 450 8C for 2 h. The XPS spectra of the samples are similar. The results show that the TiO2 structures contain not only Ti and O elements, but also a small amount of C element. The C element is ascribed to the adventitious hydrocarbons from the XPS instrument itself. The binding energies of Ti 2p at 458 eVand O 1s at 531 eVare used in our quantitative measurement. The atomic ratios of Ti:O are around 1:2, which is in good agreement with the nominal atomic composition of TiO2. Fig. 7(b) and (c) shows the high-resolution XPS spectra of the Ti 2p and the O 1s region, taken on the surface of sample E. The Ti 2p region can be tted into two peaks, the one is attributed to the Ti 2p3/2 and the other is the Ti 2p1/2, while the O 1s just one symmetrical peak. The existence of Ti 3p as shown in Fig. 7(a) is believed to cause the photocatalytic activity of TiO2. It has been reported that the photoactivity of TiO2 depends on quite a number of parameters including the crystal structure, surface area, size distribution, porosity, band gap and surface hydroxyl group density [1,21,22]. In this study,

Fig. 7. XPS spectra of sample E-1, sample A-2, sample B-3, sample G-4 (a) and the high-resolution XPS spectra of the Ti 2p (b) and the O 1s region (c) taken on the surface of sample E.

2194

X.H. Xia et al. / Materials Research Bulletin 43 (2008) 21872195

Fig. 8. Photodegradation curves of brilliant red X-3B using samples AI as photocatalysts and J-without photocatalyst.

the difference in the photocatalytic activity of the TiO2 nanocrystals should also be attributed to these aspects, especially to the structure and the surface area. In general, the anatase is taken into consideration about photoactivity, the amorphous phase titania has been known as photocatalytical inactivity, which is due to abundant defects at the surface and in the bulk promoting spontaneous recombination of photogenerated electronhole pairs with liberating energy [23]. Large surface area can help adsorbing the pollutants and is benecial to generate more electronhole pairs on the surface of the catalyst, which can degrade the absorbed pollutants sufciently. The surface areas of the samples prepared at different conditions are displayed in Table 1. Among all the as-prepared TiO2 nanostructures, sample E has the largest surface area, which allows high access of charge transfer and dye absorption. The surface areas of the samples are as follows: E > A > D > B > F > C > G > H > I, which is in well accordance with the degradations of X-3B as shown in Fig. 8. So in our study, surface area should be the key factor affecting the photocatalytic activity of the catalysts. 4. Conclusions We have successfully prepared TiO2 particles, rods, owers and sheets by hydrothermal method under acidic condition. The morphology of the samples could be controlled by adjusting the temperature, pressure and the concentration of TiCl4. Pressure is helpful for the formation of nanorods at 80 8C and nanoowers at 160 8C. It is also favorable for the crystallization of the as-prepared TiO2 nanostructures. The ower-like TiO2 has the best and the sheet-like TiO2 shows the worst photocatalytic activity. With the increase of TiCl4 concentration, the particle size increased and the surface area as well as the photocatalytic efciency of the samples decreased. The TiO2 nanostructures might also be applied to surface reaction devices such as solar cell, electro- and photochromic devices and so on. Acknowledgements This work was supported by the National Natural Science Foundation of China (No. 20207002) and the Important Item Nano-Specialized Foundation of Wuhan (No. 20041003068). References
[1] [2] [3] [4] [5] [6] J. Yu, J.C. Yu, M.K.-P. Leung, W. Ho, B. Cheng, X. Zhao, J. Zhao, J. Catal. 217 (2003) 69. J. Yu, M. Zhou, B. Cheng, X. Zhao, J. Mol. Catal. A: Chem. 246 (2006) 176. W. Guo, Z. Lin, X. Wang, G. Song, Microelectron. Eng. 66 (2003) 95. X. Zhao, J. Yu, B. Cheng, Q. Zhang, Colloids Surf. A: Physicochem. Eng. Aspects 268 (2005) 78. B. Samuneva, V. Kozhukharqv, C. Trapalis, R. Kranold, J. Mater. Sci. 28 (1993) 2353. K.S. Mazdiyasni, C.T. Lynch, J.S. Smith, J. Am. Ceram. Soc. 48 (1965) 372.

X.H. Xia et al. / Materials Research Bulletin 43 (2008) 21872195 [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] Y. Qian, Q. Chen, Z. Chen, C. Fan, G. Zhou, J. Mater. Chem. 3 (1993) 203. H. Cheng, J. Ma, L. Qi, Chem. Mater. 7 (1995) 663. S.P. Durand, J. Rouviere, C. Guizard, Colloids Surf. A: Physicochem. Eng. Aspects 98 (1995) 251. K. Ikeue, S. Nozaki, M. Ogawa, M. Anpo, Catal. Today 74 (2002) 241. Y.V. Kolenko, V.D. Maximov, A.V. Garshev, P.E. Meskin, N.N. Oleynikov, B.R. Churagulov, Chem. Phys. Lett. 388 (2004) 411. Y.V. Kolenko, B.R. Churagulov, M. Kunst, L. Mazerolles, C.C. Justin, Appl. Catal. B 54 (2004) 51. P.E. Meskon, V.K. Ivanov, A.E. Barantchikov, Ultrasono. Sonochem. 13 (2006) 47. Q.H. Zhang, L. Gao, J.K. Guo, J. Inorg. Mater. 15 (2000) 21. Y. Li, J. Liu, Z. Jia, Mater. Lett. 60 (2006) 1753. J. Livage, M. Henry, C. Sanchez, Prog. Solid State Chem. 18 (1988) 259. Y. Li, T.J. White, S.H. Lin, J. Solid State Chem. 17 (2004) 1372. J.P. Liu, X.T. Huang, J.X. Duan, H.H. Ai, P.H. Tu, Mater. Lett. 59 (2005) 3710. Y. Xie, C. Yuan, X. Li, Colloids Surf. A: Physicochem. Eng. Aspects 252 (2005) 87. L. Fan, F. Yang, W. Yang, Sep. Purif. Technol. 34 (2004) 89. A. Narayanasamy, V.A. Maroni, R.W. Siegel, J. Mater. Res. 4 (1989) 1246. L. Shi, C. Li, H. Gu, D. Fang, Mater. Chem. Phys. 62 (2000) 62. Y. Xie, C. Yuan, Appl. Catal. B: Environ. 46 (2003) 251.

2195

Das könnte Ihnen auch gefallen