Vous êtes sur la page 1sur 5

Journal of Materials Chemistry

Cite this: J. Mater. Chem., 2011, 21, 4888 www.rsc.org/materials

Dynamic Article Links <

View Online

PAPER

Mesoporous hollow TiO2 microspheres with enhanced photoluminescence prepared by a smart amino acid template
Shangjun Ding,a Fuqiang Huang,*a Xinliang Mou,b Jianjun Wua and Xujie L ua
Downloaded by Amity University on 12 August 2011 Published on 18 February 2011 on http://pubs.rsc.org | doi:10.1039/C0JM03628E

Received 25th October 2010, Accepted 12th January 2011 DOI: 10.1039/c0jm03628e The simplest amino acid, glycine (C2H5NO2), was used as a multifunctional smart template to prepare mesoporous hollow TiO2 microspheres via a facile solvothermal method. In the fabrication process, glycine serves concurrently as a sacricial template, complexing agent and water supplier. The hollow spheres were characterized as TiO2/organic hybrid structures and converted into well-dened mesoporous structures by calcination. The growth mechanism of the hollow spheres was demonstrated to include the self-conglobated smart template above the softening temperature, a consequent in situ hydrolysis on the template surface and then a sacricial core evacuation process. From the demonstrated mechanism, the cavity volume can be easily controlled by the mass of crude glycine used. This novel method has successfully been applied in the synthesis of Eu-doped TiO2 hollow spheres. The Eu-doped sample exhibits stronger photoluminescence owing to the unique microstructures, compared to conventional hydrothermal samples.

1. Introduction
Titanium dioxide (TiO2) has excellent chemicophysical properties and unique applications in photocatalysis, gas sensors, solar cells, luminescence lms and Li-ion batteries, etc.17 Extensive research studies have been focused on controlling the microstructure and morphology of TiO2 to achieve novel and enhanced properties. Typically, TiO2 hollow spheres attract great interest due to their low density, high surface area, delivering ability, surface permeability, light-trapping effect, etc.810 TiO2 hollow spheres can be prepared with the assistance of various templates (polystyrene,11 carbon beads,12 micelle and emulsion droplets,13 and bubbles14) and an appropriate surfactant. The latter material has to ensure that the titanium precursor hydrolyzes only on the surface of the template. The preparation of TiO2 hollow spheres currently requires a simpler process, lower cost and better environmental compatibility. Amino acids, as typical biomolecules with polyfunctional groups, have a low softening point, low solubility in alcohols, selective adsorptivities and particular metal cation complexations.15 Based on these unique physical and chemical properties, we propose a solid amino acid to serve concurrently as a smart template, surfactant and reactant in the preparation of hollow

structures. Amino acid particles in the reaction conditions can be converted into micron-sized spheres due to their low softening point and low solubility in alcohols. The obtained amino acid spheres can act as sacricial templates to react with alcohols to generate water and in situ-hydrolyze the titanium precursor on the surface to grow mesoporous hollow TiO2 structures. Herein, as an initial attempt, we have developed a novel solvothermal approach to fabricate mesoporous TiO2 hollow spheres with the assistance of glycine. This simplest of amino acids plays a multifunctional role in our synthesis, despite its simple molecular structure (C2H5NO2). Hollow TiO2 spheres were obtained by a proposed growth mechanism. This synthetic route tremendously simplies the experimental procedure, which was also applied in the fabrication of Eu-doped TiO2 hollow spheres. The luminescence of these Eu-doped hollow spheres was investigated. The enhanced luminescent property was explained by the unique microstructure of the phosphor materials.

2. Experimental
2.1 Synthesis All the chemicals (analytical grade) were commercially obtained without further purication. In a typical synthesis, a suspension of 0.6 g glycine (C2H5NO2) and 30 mL anhydrous ethanol was stirred in a sealed vessel for 1 h at room temperature. 1.5 mL titanium n-butoxide (Ti(OBu)4, TNB) was added to the suspension, stirred for another 1 h and transferred into a 50 mL Teon-lined stainless steel autoclave. The steel autoclave was heated at 200  C for 20 h without stirring and then cooled gradually to room temperature. The resulting precipitate was
This journal is The Royal Society of Chemistry 2011

a CAS Key Laboratory of Materials for Energy Conversion, Shanghai Institute of Ceramics, Chinese Academy of Sciences (SICCAS), Shanghai, 200050, P. R. China. E-mail: huangfq@mail.sic.ac.cn; Fax: +86 21 5241 6360 b Inorganic Materials Analysis and Testing Center, SICCAS, Shanghai, 200050, P. R. China Electronic supplementary information (ESI) available: See DOI: 10.1039/c0jm03628e

4888 | J. Mater. Chem., 2011, 21, 48884892

View Online

collected, washed with anhydrous ethanol three times and dried at 60  C overnight prior to being characterized. In the preparation of the Eu-doped TiO2 sample, anhydrous europium nitrate was initially introduced into the solvent without changing any other experimental conditions. For the calcination treatment, the as-prepared sample was heated in air at 450  C for 5 h. 2.2 Characterization The crystalline structure and morphology of the samples were characterized by X-ray diffraction measurements (XRD, Bruker D8 ADVANCE, Cu-Ka radiation), scanning electron microscopy (SEM, JSM 6510) and transmission electron microscopy (TEM, JEM 2100F), respectively. The chemical composition of the samples were determined by energy dispersive X-ray spectroscopy (EDS) attached to the TEM. UV-vis diffusing reectance spectra were measured at room temperature on a spectrophotometer (Hitachi U3010). FTIR information was recorded using a Shimadzu IR Prestige-21 instrument in the range 4004000 cm1. Room temperature photoluminescence (PL) spectra were recorded on a spectrouorimeter (FluoroMax 4, HORIBA Jobin Yvon) using an Xe lamp as the excitation source. Thermal analyses were carried out on a simultaneous TG-DTA/DSC apparatus (STA409PC Luxx, Neutsch) with a heating rate of 10  C min1 in owing air. Nitrogen adsorption desorption isotherms at 77 K were measured on a Micromeritics Tristar 3000 system. Pore size distributions were calculated from desorption branches of isotherms by the BarrettJoyner Halenda (BJH) method.

3. Results and discussion


3.1 Morphology and structure of the TiO2 hollow spheres The as-prepared samples were TiO2/organic hybrid structures, as evidenced from the XRD patterns and UV-vis absorption spectra before and after calcination, as shown in Fig. 1a and 1b, respectively. All the XRD diffraction peaks can be assigned to

anatase TiO2 (JCPDS# 21-1272). For the as-prepared sample before calcination treatment, the obvious broadened diffraction peaks indicate a crystallite size of 4.8 nm, calculated from Scherrer equation, based on the main diffraction peak (101). The UV-vis absorption spectrum exhibits a strong shoulder absorption from 360 to 800 nm due to light-absorbing organics in the TiO2 matrix, consistent with the yellowish color of the asprepared sample. After the calcination treatment, the full-width at half-maximum (FWHM) of the respective diffraction peak decreases (Fig. 1a), and the crystallinity of the sample is improved. The absorption edge of the white-colored sample starts at 395 nm, as shown in Fig. 1b. It is believed that the hybrid organic species are burned off during the calcination treatment. The TiO2/organic hybrid structure was conrmed by its FTIR spectrum and TG/DSC measurements. As shown in Fig. 1c, the organic FTIR absorption peaks in spectrum A were observed to be at 1685 cm1 (C]O stretching vibration), 1470 cm1 (CH bending vibration) and 1338 cm1 (CN stretching vibration), but they disappeared after the calcination treatment (spectrum B). In the TG/DSC curves (Fig. 1d), there are two major weight loss stages, and the relevant DSC peaks are close to 110 and 340  C. The rst weight loss stage is caused by the evaporation of physically absorbed water and the second loss, along with the exothermic nature in the DSC curve, suggests some organic (peptide) combustion. The elected temperature 450  C was high enough to burn off the hybrid organics, as shown in Fig. 1d. The irregular morphology of the raw glycine added into the ethanol solvent is shown in Fig. 2a. However, the as-prepared hybrid TiO2 sample via this glycine-assisted solvothermal reaction has a well-dened hollow microsphere structure, as illustrated in Fig. 2b2d. TiO2 microspheres are about 17 mm in diameter with a particularly smooth surface; a broken sphere with a hollow interior is shown in Fig. 2d. Interestingly, the spherical particles can stick together to form microsphere dimmers and chains (Fig. 2b). The characteristic hollow structures are further conrmed by the corresponding TEM micrographs in Fig. 3a and 3b. The wall thickness of the hollow microspheres were identied to be of several hundred

Downloaded by Amity University on 12 August 2011 Published on 18 February 2011 on http://pubs.rsc.org | doi:10.1039/C0JM03628E

Fig. 1 (a) XRD pattern, (b) UV-vis absorption spectra and corresponding photographs (inset), (c) FTIR spectra of the pre- (curve A) and post-calcination (curve B) TiO2 hollow sphere samples; (d) TG/DSC curves for the as-prepared sample.

Fig. 2 (a) SEM image of the crude glycine particles and SEM images of the as-prepared TiO2 spheres at (b) low magnication, (c) high magnication and (d) showing a broken microsphere.

This journal is The Royal Society of Chemistry 2011

J. Mater. Chem., 2011, 21, 48884892 | 4889

View Online

in the pore size distribution curve, as calculated by the BJH method. 3.2 Self-template effect and cavity size control The template effect of glycine and the size controllability of our synthetic procedure can be proven by sorting the mass of the crude glycine. In the mass-sorting process, a suspension of glycine and EtOH was precipitated for 20 min after stirred for 1 h; the large glycine particles at the bottom were ltered off and the relatively stable suspension was transferred to an autoclave for the solvothermal reaction. The consequent TiO2 sphere morphologies without and with mass-sorting are shown in Fig. 5a and 5b, respectively. The latter sample clearly consists of homogenous particles and has a narrow particle size distribution. The corresponding statistical histograms of the particle size distributions are displayed in Fig. 5c and 5d. The mass-unsorted sample has two distribution peaks centered at 2.5 and 5.7 mm, and the mass-sorted sample only has one peak at 2.5 mm. This indicates that the mass of the crude glycine particles determines the spherical template size to control the size of the cavity. The impetus to form a spherical morphology is generally attributed to a surface or interface effect. In our case, the crude glycine particles (Fig. 2a) hardly dissolve in absolute alcohol. When the solvothermal reaction temperature is up to 200  C, the irregular glycine particles are turned into spherical shapes as a result of their low softening point (180  C) and interface tension. Afterwards, the spherical quasi-liquid droplets of glycine act as sacricial templates to grow TiO2 hollow spheres. This self-conglobation process indicates that the irregular hard templates of the amino acid are feasible for forming a spherical morphology. Therefore, it is more facile to prepare the templates in this smart glycine-assisted process compared to other traditional template-based methods. 3.3 Growth mechanism of the TiO2 hollow microspheres The formation of TiO2 in anhydrous ethanol relies on the hydrolysation of titanium n-butoxide with the aid of water. Our

Downloaded by Amity University on 12 August 2011 Published on 18 February 2011 on http://pubs.rsc.org | doi:10.1039/C0JM03628E

Fig. 3 (a,b) TEM images, (c) the HRTEM image and (d) the SAED pattern of the as-prepared TiO2 hollow microspheres.

nanometers. The inner cavities of some of the spheres are connected with each other. The high-resolution TEM (HRTEM) image (Fig. 3c) shows that the microspheres are composed of nanocrystals with a size of about 5 nm. The polycrystalline anatase phase is demonstrated by the selected area electron diffraction (SAED) pattern (Fig. 3d), which is in accordance with the XRD powder pattern. The TiO2/organic hybrid hollow spheres can be converted into mesoporous hollow structures by calcination at 450  C. The BET surface area of the mesoporous sample is 26.98 m2 g1. The nitrogen adsorptiondesorption isotherm of the post-heat treatment sample is shown in Fig. 4. The desorption branch exhibits two major capillary condensation steps, appearing in the relative pressure ranges of 0.40.6 and 0.70.85. These steps correspond to two distinct peaks at 3.5 and 6.1 nm, respectively,

Fig. 4 The nitrogen adsorptiondesorption isotherm and corresponding pore size distribution (inset) of post-treatment TiO2 hollow spheres.

Fig. 5 SEM images and diameter distribution histograms of TiO2 hollow spheres (a,c) without and (b,d) with size-sorting processing.

4890 | J. Mater. Chem., 2011, 21, 48884892

This journal is The Royal Society of Chemistry 2011

View Online

proposed method does not introduce any water to the hydrolysation of the titanium alkoxide. Without the presence of glycine in the solvothermal reaction, no TiO2 precipitate was observed. Therefore, it must be glycine that indirectly supplies water for the hydrolysis. The molecule glycine (C2H5NO2) contains two functional groups, an NH2 (amino group) and a COOH (carboxyl group). Both the self-reaction of glycine (eqn (1)) and the esterication reaction of glycine and ethanol (eqn (2)) are able to produce water16 to cause the hydrolysation. NH2CH2COOH + NH2CH2COOH / NH2CH2CONHCH2COOH + H2O
Downloaded by Amity University on 12 August 2011 Published on 18 February 2011 on http://pubs.rsc.org | doi:10.1039/C0JM03628E

and the dissolution of more soluble products, such as ester formed by the condensation reaction. 3.4 Photoluminescence of Eu-doped TiO2 hollow spheres The glycine-assisted smart template method is a very simple way of fabricating mesoporous TiO2 hollow spheres due to in situ hydrolysis and a self-templating effect. Based on this method, doped TiO2 hollow spheres can be synthesized by introducing other alcohol-soluble chemicals into the reaction, such as metal alkoxides or anhydrous nitrates. Herein, by adding europium nitrate into the precursor solution, Eu-doped TiO2 hollow spheres with an intense photoluminescence were easily obtained. Typical excitation and emission spectra of the Eu-doped TiO2 hollow spheres are depicted in Fig. 7a. The strongest emission peak, centered at 613 nm, is attributed to the Eu3+ 4f4f transition from 5D0 / 7F2. By monitoring the emission line at 613 nm, the excitation lines at 394, 465 and 534 nm can be ascribed to the ff inner-shell transitions within the electronic conguration of Eu3+ 4f6.17 The Eu-doped powder displayed strong red light luminescence under a 365 nm ultraviolet lamp, as shown in the inset of Fig. 7a. Apparently, the powders can be also be excited by 465 nm blue light (Fig. 7b), which is in accordance with the emission of blue LED chips. Therefore, the Eu-doped powder is also a blue LED-driven phosphor. The existence of Eu in this sample was conrmed by the EDS spectrum in Fig. S3 of the ESI. The TEM images and XRD pattern (ESI, Fig. S3) indicate that the crystallinity of the hollow spheres decreased due to Eu3+ doping. In order to study the inuence of the microstructure on luminescence by using the different synthetic methods, the Eu-doped TiO2 phosphor was prepared by the conventional hydrothermal method. For comparison, the reaction conditions and the microstructure of the nanoparticles are shown in Table S1 and Fig. S4 of the ESI. Under the excitation of 465 nm blue light, both Eu-doped samples can emit red light, as shown in their PL spectra presented in Fig. 7b. The PL intensity of the TiO2 hollow spheres is about 2.5 times larger than that of common hydrothermal nanoparticles. Eu doping into the TiO2 lattice is very hard due to the  and large difference between the atomic radii of Ti4+ (0.745 A)  In the present study, the relatively slow Eu3+ (1.087 A). in situ hydrolysis makes Eu3+ doping easier and more dispersive. The mesoporous hollow sphere assembled by nanocrystals with a size of about 5 nm has a higher abundance of grain boundaries than the nanoparticle sample. The grain boundary region can be

(1)

NH2CH2COOH + CH3CH2OH / NH2CH2COOCH2CH3 + H2O (2) In order to distinguish the relationship between the water generation route and the as-obtained TiO2 morphology, n-octane was used to replace ethanol in the reaction. However, only irregular anatase TiO2 particles were obtained, as shown in Fig. S1 of the ESI. By using other alcohol solvents, such as n-butanol and isopropanol, the spherical morphology of the TiO2 particles remained (ESI, Fig. S2). The esterication reaction, occurring at the interface between glycine and the alcohol, plays a more important role in the formation of the hollow spherical morphology than does the self-reaction of glycine. A mechanism is proposed for the fabrication of TiO2 hollow spheres with the aid of smart templates of glycine, as shown in Fig. 6. The irregular crude glycine particle is rstly transformed into a spherical template by interface tension in the reaction conditions above its softening temperature. The template surface contains a large number of dangling amine and carboxyl groups in order to exchange ligands with the titanium alkoxide.15 This effect causes the titanium precursor to gather near the template, avoids the adding of any other additional surfactants and ensures hydrolysis on the templates surface. The interface esterication reaction generates water for the in situ hydrolysis of TNB. The initially formed nano-TiO2 grains are assembled on the surface of spherical glycine to construct the spherical morphology. The TiO2 shell is gradually thickened due to sustainable hydrolysis, and the interior cavity is formed by a core evacuation process through the integration of organics into the nal TiO2 spheres

Fig. 6 A scheme showing the formation process of the TiO2 hollow spheres: (A) surface tension effect, (B) in situ hydrolysis, (C) shell formation, (D) shell growth, (E) core evacuation, (F) the stickingtogether process, and (G) shell growth and core evacuation processes.

Fig. 7 (a) The excitation spectrum (left), emission spectrum excited at 394 nm (right) and a photograph of the Eu-doped TiO2 hollow spheres under a 365 nm ultraviolet lamp (inset). (b) PL spectra of the hollow spheres and nanoparticles excited at 465 nm.

This journal is The Royal Society of Chemistry 2011

J. Mater. Chem., 2011, 21, 48884892 | 4891

View Online

benecial to accommodate Eu3+ ions to suppress the EuEu interaction.18 Therefore, the highly bright red uorescence of the Eu-doped TiO2 hollow spheres is ascribed to their unique microstructure, obtained via this glycine-assisted method.

4 Conclusions
A novel and convenient solvothermal approach has been developed to prepare mesoporous hollow TiO2 microspheres with the assistance of a smart template, glycine. The as-prepared hollow spheres were found to be TiO2/organic hybrid structures, and were converted into mesoporous structures after heat treatment at 450  C, as conrmed by UV-vis absorption and FTIR transmission spectra, and thermoanalysis. Eu-doped TiO2 hollow spheres prepared by the same method exhibited an enhanced photoluminescence due to their unique microstructure. A low softening point, slight solubility in alcohol, in situ hydrolysis, a self-template effect and specic coordination activity make the solid glycine particles become smart templates. Glycine in this simple synthetic process acts as a sacricial template, complexing agent and water supplier. The study provides enlightenment into the design of other hollow structures by using amino acids as multifunctional templates.

Acknowledgements
This work was nancially supported by National 973 Program of China Grant no. 2007CB936704 and 2009CB939903, National Science Foundation of China Grant no. 50772123 and 50821004, and the Science and Technology Commission of Shanghai Grant no. 0952nm06500 and 08JC1420200.

References
, L. L. Zhang, F. Q. Huang and F. F. Xu, Eur. J. 1 J. J. Wu, X. J. Lu Inorg. Chem., 2009, 2789. 2 M. Li, Z. L. Hong, Y. N. Fang and F. Q. Huang, Mater. Res. Bull., 2008, 43, 2179.

3 M. Ferroni, V. Guidi, G. Martinelli, G. Faglia, P. Nelli and G. Sberveglieri, Nanostruct. Mater., 1996, 7, 709. 4 S. C. Yang, D. J. Yang, J. Kim, J. M. Hong, H. G. Kim, I. D. Kim and H. Lee, Adv. Mater., 2008, 20, 1059. , X. L. Mou, J. J. Wu, D. W. Zhang, L. L. Zhang, F. Huang, 5 X. J. Lu F. F. Xu and S. M. Huang, Adv. Funct. Mater., 2010, 20, 509. 6 M. Luo, K. Cheng, W. J. Weng, C. L. Song, P. Y. Du, G. Shen and G. R. Han, Nanoscale Res. Lett., 2009, 4, 809. 7 J. P. Wang, Y. Bai, M. Y. Wu, J. Yin and W. F. Zhang, J. Power Sources, 2009, 191, 614. 8 H. X. Li, Z. F. Bian, J. Zhu, D. Q. Die, G. S. Li, Y. N. Huo, H. Li and Y. F. Lu, J. Am. Chem. Soc., 2007, 129, 8406. 9 S.-H. A. Lee, N. M. Abrams, P. G. Hoertz, G. D. Barber, L. I. Halaoui and T. E. Mallouk, J. Phys. Chem. B, 2008, 112, 14415. 10 U. Jeong, S. H. Im, P. H. C. Camargo, J. H. Kim and Y. Xia, Langmuir, 2007, 23, 10968. 11 Y. Kondo, H. Yoshikawa, K. Awaga, M. Murayama, T. Mori, K. Sunada, S. Bandow and S. Iijima, Langmuir, 2008, 24, 547; G. Q. Li, C. Y. Liu and Y. Liu, J. Am. Ceram. Soc., 2007, 90(8), 2667; X. F. Wu, Y. J. Tian, Y. B. Cui, L. Q. Wei, Q. Wang and Y. F. Chen, J. Phys. Chem. C, 2007, 111, 9704. 12 Y. H. Ao, J. J. Xua, D. G. Fu and C. W. Yuan, J. Hazard. Mater., 2009, 167, 413; H. Q. Wang, Z. B. Wu and Y. Liu, J. Phys. Chem. C, 2009, 113, 13317. 13 R. Inaba, T. Fukahori, M. Hamamoto and T. Ohno, J. Mol. Catal. A: Chem., 2006, 260, 247; Y. D. Wang, A. N. Zhou and Z. Y. Yang, Mater. Lett., 2008, 62, 1930. 14 X. X. Li, Y. J. Xiong, Z. Q. Li and Y. Xie, Inorg. Chem., 2006, 45, 3493; L. Jiang, Y. J. Zhong and G. C. Li, Mater. Res. Bull., 2009, 44, 999. 15 K. Kanie and T. Sugimoto, Chem. Commun., 2004, 1584; Y. Wang, Q. S. Zhu and H. G. Zhang, Chem. Commun., 2005, 5231; N. Ikawa, T. Kimura, Y. Oumi and T. Sano, J. Mater. Chem., 2009, 19, 4906; V. Prevot, N. Caperaa, C. Taviot-Gu eho and C. Forano, Cryst. Growth Des., 2009, 9, 3646. 16 C. Wang, Z. X. Deng and Y. D. Li, Inorg. Chem., 2001, 40, 5210. 17 H. You and M. Nogami, J. Phys. Chem. B, 2004, 108, 12003; X. M. Liu, C. X. Li, Z. Y. Cheng and J. Lin, J. Phys. Chem. C, 2007, 111, 16007. n and 18 A. Conde-Gallardo, M. Garc a-Rocha, I. Hem andez-Caldero R. Palomino-Merino, Appl. Phys. Lett., 2001, 78, 3436; J. B. Yin, L. Q. Xiang and X. P. Zhao, Appl. Phys. Lett., 2007, 90, 113112; L. Li, C. K. Tsung, Z. Yang, G. D. Stucky, L. D. Sun, J. F. Wang and C. H. Yan, Adv. Mater., 2008, 20, 903.

Downloaded by Amity University on 12 August 2011 Published on 18 February 2011 on http://pubs.rsc.org | doi:10.1039/C0JM03628E

4892 | J. Mater. Chem., 2011, 21, 48884892

This journal is The Royal Society of Chemistry 2011

Vous aimerez peut-être aussi