Вы находитесь на странице: 1из 7

Subscriber access provided by EAST CHINA UNIV OF SCI & TECH

Article

Preparation, Photocatalytic Activity, and Mechanism of Nano-TiO Co-Doped with Nitrogen and Iron (III)
2

Ye Cong, Jinlong Zhang, Feng Chen, Masakazu Anpo, and Dannong He


J. Phys. Chem. C, 2007, 111 (28), 10618-10623 DOI: 10.1021/jp0727493 Publication Date (Web): 20 June 2007 Downloaded from http://pubs.acs.org on April 22, 2009

More About This Article


Additional resources and features associated with this article are available within the HTML version: Supporting Information Links to the 4 articles that cite this article, as of the time of this article download Access to high resolution figures Links to articles and content related to this article Copyright permission to reproduce figures and/or text from this article

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036

10618

J. Phys. Chem. C 2007, 111, 10618-10623

Preparation, Photocatalytic Activity, and Mechanism of Nano-TiO2 Co-Doped with Nitrogen and Iron (III)
Ye Cong, Jinlong Zhang,*, Feng Chen, Masakazu Anpo, and Dannong He
Lab for AdVanced Materials and Institute of Fine Chemicals, East China UniVersity of Science and Technology, 130 Meilong Road, Shanghai 200237, China, Department of Applied Chemistry, Graduate School of Engineering, Osaka Prefecture UniVersity, 1-1 Gakuen-cho, Sakai, Osaka 599-8531, Japan, and Shanghai National Engineering Research Center for Nanotechnology, 245 Jiachuan Road, Shanghai 200237, China ReceiVed: April 9, 2007; In Final Form: May 14, 2007

Nanoparticles of titanium dioxide co-doped with nitrogen and iron (III) were first prepared using the homogeneous precipitation-hydrothermal method. The structure and properties of the co-doped were studied by XRD, XPS, Raman, FL, and UV-diffuse reflectance spectra. By analyzing the structures and photocatalytic activities of the undoped and nitrogen and/or Fe3+-doped TiO2 under ultraviolet and visible light irradiation, the probable mechanism of co-doped particles was investigated. It is presumed that the nitrogen and Fe3+ ion doping induced the formation of new states closed to the valence band and conduction band, respectively. The co-operation of the nitrogen and Fe3+ ion leads to the much narrowing of the band gap and greatly improves the photocatalytic activity in the visible light region. Meanwhile, the co-doping can also promote the separation of the photogenerated electrons and holes to accelerate the transmission of photocurrent carrier. The photocatalyst co-doped with nitrogen and 0.5% Fe3+ shows the best photocatalytic activity, the degradation efficiencies of which were improved by 75% and 5% under visible and ultraviolet irradiation, respectively, compared with the pure titania.

1. Introduction One of the most intriguing application areas of titanium dioxide is the photocatalytic activity for environment protection since it can decompose a large number of organic and inorganic pollutants.1-3 However, because of the wide band gap of titanium dioxide, its practical application is inhibited for the low photon utilization efficiency and need of an ultraviolet excitation source which accounts for only small fraction of the solar light (3-5%). Therefore, it is an important and challenging issue to develop new TiO2 photocatalytic system with enhanced activities under both UV and visible light irradiation compared with the bare TiO2, improving the utilization efficiency of the solar energy. Impurity doping is one of the typical approaches to extend the spectral response of the titanium dioxide to visible light region. Some metal elements such as Fe, Cr, Co, Mo, and V have been employed to tune the electronic structure and enhanced the photocatalytic activity of the titanium dioxide.4-8 It has been hypothesized that the advantages of the doping of the metal ion in TiO2 are the temporary trapping of the photogenerated charge carriers by the dopants and the inhibition of their recombination during migration from inside of the material to the surface or the enhanced adsorption of the goal pollutants onto the doping ion surface sites. Choi et al. have reported that TiO2, doped with 21 kinds of metal ions prepared by sol-gel method, exhibited various colors depending on the
* Corresponding author. Tel: +86-21-64252062. Fax: +86-21-64252062. E-mail: jlzhang@ecust.edu.cn. East China University of Science and Technology. Osaka Prefecture University. Shanghai National Engineering Research Center for Nanotechnology.

ion used.9 Anpo and co-worders suggested that the injected metal ions isomorphicaly replaced Ti4+ ions in TiO2 and decreased the band gap to enable absorb the visible light.10 Recently, many efforts have been attempted by modification titanium dioxide with nonmetals, such as B, C, N, S, and F, so as to efficiently extend photoresponse from UV to visible light region.11-16 Furthermore, some theoretical calculations have also been performed to suggest that anion doping of TiO2 has considerable effects on the band gap alteration.13,17 The success in nitrogen doping and increasing the photocatalytic activity of TiO2 in visible light region provides good opportunities for extensive applications such as oxidation of CO, ethanol, gaseous 2-propanol, acetaldehyde, NOx and the decomposition of dyes such as methylene blue. It is well-known that nitrogen doping can improve the phtocatalytic activity of titania in the visible region, while the UV photocatalytic activities are seldom studied. Unfortunately, in many cases, the UV photocatalytic activities are relatively low. Therefore, it is a challenging issue to develop a photocatalyst with higher photocatalytic activity under both ultraviolet and visible light irradiation. To the best of our knowledge, few reports are published on TiO2 photocatalyst co-doped with nitrogen and iron (III) for photo-oxidative degradation of organic pollutants. Here, we prepare nanocrystalline TiO2 photocatalysts co-doped with nitrogen and iron (III) by homogeneous precipitation-hydrothermal method. This method can avoid the agglomeration and growth of particles in the process of higher temperature calcinations. The prepared photocatalysts manifest higher photocatalytic activities for photodegradation of rhodamine B under both ultraviolet and visible light irradiation. The factors influencing the photocatalytic activity and the probable mechanism of co-doped nanoTiO2 are investigated.

10.1021/jp0727493 CCC: $37.00 2007 American Chemical Society Published on Web 06/20/2007

Nano-TiO2 Co-Doped with Nitrogen and Iron (III) 2. Experimental Section 2.1. Preparation of Samples. Tetrabutyltitanate and ferric nitrate were used as starting materials. Ammonium chloride was used as a nitrogen source. A desired amount of solid ferric nitrate and 17.0 mL tetrabutyltitanate was mixed with 25 mL absolute alcohol. The mixture was slowly dropped into a solution of ammonium chloride in water with continuously stirring. The formed gel precipitation was aging for 24 h to completely hydrolyze. Then the obtained mixture was changed into 100 mL Teflonlined stainless steel autoclave and then heated under 180 C for 12 h. After the autoclave was cooled to room temperature, the precipitate at the bottom of the autoclave was obtained and washed with ethanol and deionized water several times. The product was kept in infrared drying oven to dry overnight and the nitrogen and Fe3+-co-doped TiO2 photocatalysts were obtained. The Fe3+-doping concentration (x) was chosen as 0.1, 0.5, 1.0, 1.5, and 2.0, which was the mole percentage of Fe3+ in the theoretical titania powder. The obtained corresponding concentration photocatalysts were denoted as xFe-TiO2 and N- xFe-TiO2. 2.2. Characterization Methods. X-ray diffraction (XRD) patterns of all samples were collected in the range 20-80 (2) using Rigaku D/MAX 2550 diffractometer (Cu KR radiation, ) 1.5406 ), operated at 40 kV and 100 mA. The crystallite size was estimated by applying the Scherrer equation to the fwhm of the (101) peak of anatase and the (110) peak of rutile, with R-silicon (99.9999%) as a standard for the instrumental line broading. The Raman spectra were recorded on an inVia + Reflex spectrometer equipped with an optical microscope at room temperature. For excitation, the 514.5 nm line from an Ar+ ion laser (Spectra Physics) was focused, with an analyzing spot about 1 m, on the sample under the microscope. The instrument employed for XPS studies was Perkin-Elmer PHI 5000C ESCA System with Al KR radiation operated at 250 W. The shift of binding energy due to relative surface charging was corrected using the C1s level at 285.0 eV as an internal standard. The electron paramagnetic resonance (EPR) spectra were recorded at 298 K using a Bruker AVANCE 500 electron paramagnetic resonance spectrometer. The UV-visible absorbance spectra were obtained for the dry-pressed disk samples using a Scan UV-visible spectrophotometer (Varian, Cary 500) equipped with an integrating sphere assembly, using BaSO4 as reflectance sample. The photoluminescence (PL) spectra were measured through fluorospectrophotometer (Cary Eclipse) by using 280 nm line of Xe lamp as excitation source at room temperature. 2.3. Photocatalytic Activity Measurements. The photocatalytic activity was evaluated by measuring the decomposition of the distilled water solution of rhodamine B (with a concentration of 20 mg/L) under UV and visible light irradiation. UV irradiation was carried out using a 300 W high-pressure mercury lamp, the irradiation wavelength of which was predominated at 365 nm. A 1000 W halogen lamp was used as the visible light source of the homemade photoreactor, surrounded with cool water at the outer wall through a quartz jacket. The shortwavelength components (< 420 nm) of the light were cut off using a glass optical filter. For a typical photocatalytic experiment, a total of 0.05 g catalyst powders were added into 50 mL of the above rhodamine B solution in quartz tube. Prior to irradiation, the suspensions were magnetically stirred in dark for 30 min to ensure the establishment of an adsorption/ desorption equilibrium. The above suspensions were kept under constant air-equilibrated conditions before and during the irradiation. At given time intervals, about 4 mL aliquots were

J. Phys. Chem. C, Vol. 111, No. 28, 2007 10619

Figure 1. X-ray diffraction patterns of xFe-TiO2 (a) and N-xFeTiO2 (b) with different Fe3+-doping concentration: (a, a) 0; (b, b) 0.05%; (c, c) 0.1%; (d, d) 0.5%; (e, e) 1.0%; (f, f) 1.5%; (g, g) 2.0%.

sampled, centrifuged, and filtered through a 0.22m membrane filter to remove the particles. The filtrates were analyzed by recording variations in the absorption in UV-visible spectra of rhodamine B using a Cary 100 Ultraviolet-visible spectrometer. In order to diminishing the experimental error, we repeated experiment at least three times for the same sample and chose the mean value in our experiment. Moreover, the photocatalytic tests of xFe-TiO2 and N-xFe-TiO2 for the same Fe 3+-doping concentration are parallel determined at the same time and conditions. 3. Results and Discussion 3.1. Structures and Properties of Photocatalysts. In Figure 1, the XRD patterns of xFe-TiO2 and N-xFe-TiO2 with different Fe3+-doping concentration are provided. The major crystalline phase detected in all samples is anatase. It can be seen from Figure 1 b that the samples after doping with nitrogen show trace amount of brookite phase. Given the ammonium chloride was used as the nitrogen source, this response is reasonable because a high chloride concentration promotes the crystallization of brookite.18 It is noteworthy that no iron oxide or FexTiOy peak could be observed in the XRD spectra. We think all the iron ions were incorporated into the structures of titania and replaced titanium ion or located at interstitial site. The crystallite sizes of samples were calculated using the Debye-Scherrer equation and the values are given in Table 1. The crystallite size of all prepared samples is about 10-12 nm, which indicates the post hydrothermal can effectively prompt

10620 J. Phys. Chem. C, Vol. 111, No. 28, 2007


TABLE 1: The Diameter of the Crystallites of Anantase Phase for xFe-TiO2 and N-xFe-TiO2 with Different Fe3+-Doping Concentration
no. 1 2 3 4 5 6 7 x (Fe mol %) 0.00 0.05 0.10 0.50 1.00 1.50 2.00 xFe-TiO2 (nm) 10.26 10.85 11.59 11.23 11.18 15.14 12.15 N-xFe-TiO2 (nm) 9.63 9.98 10.52 9.80 9.82 10.59 10.44

Cong et al.

crystallization and inhibit the grain growth compared with the high temperature annealing. Furthermore, the crystallite size of the samples after nitrogen doping is smaller than that of the undoped samples and corresponding concentration Fe3+-doped samples. It is thought that nitrogen doping prolongs the crystallization of anatase and retards the transformation of amorphous titania to anatase since the crystallinity of nitrogen doped samples is lower than that of undoped samples. Raman spectroscopy is a powerful technique for the investigation of various phases of titanium dioxide. This technique is capable of elucidating the photocatalyst structural complexity as peaks from each material are clearly separated in frequency, and therefore, the phases are easily distinguishable. The modes A1g (519 cm-1), B1g (399 cm-1), and Eg (144, 197, and 639 cm-1) are Raman-active, and thus, six fundamental transitions are expected in the Raman spectrum of anatase.19 Figure 2 shows the Raman spectra of different samples. Raman peak at about 146 cm-1 is observed for all the samples, which is attributed to the main Eg anatase vibration mode. Moreover, vibration peaks at 199 cm-1 (Eg, weak), 399 cm-1 (B1g), 516 cm-1 (A1g), and 640 cm-1 (Eg) are present in the spectra for all samples, which indicates that anatase TiO2 nanoparticles are the predominant species. Furthermore, there is no peak attributed to iron oxide observed, which is consistent with the results of XRD patterns. From a measurement of the maximum of the low-frequency Raman band, it is possible to determine the nanoparticle size since the particle size can cause large shifts in the location of the scattered Raman peaks and their widths, namely the quantum size confinement effect.20,21 Consequently, compared the lowest frequency peak at 146 cm-1 of the different samples, it can be evidently seen that the intensities of this peak are dramatically increased and its widths are broadened after nitrogen doping. It indicates that the crystallinity is enhanced and the particle size is decreased, which is corresponding to the XRD results. Up to now, the assignment of the XPS peak of N1s has still been under debate, and controversial hypotheses have been

Figure 3. The XPS spectra for the N1s region of N-TiO2 and N-0.5%Fe-TiO2.

Figure 2. Raman spectra of different samples.

provided. Since the preparation methods and conditions largely affect nitrogen XPS spectral features, the peak position may be different from the literatures. In addition, the nitrogen source is different, which could probably influence the characteristics of the nitrogen state. By analyzing the preparation methods and peak characterizations in the XPS spectra of the nitrogen-doped titania, we think that the peak position lies in the range of 397401 eV is much reasonable to the one when using the wet chemical method. In Figure 3, the XPS spectra for the N 1s region of nitrogen doped and nitrogen and iron codoped titania are presented. The main peak appears at about 399 eV. This peak can be assigned to substitutional nitrogen replaced one O atom of the TiO2 lattice or the interstitial nitrogen bound to one lattice oxygen and formed a NO species.22 Both the substitutional and interstitial nitrogen could affect the electronic band structure of TiO2 and improve the photocatalytic activity in the visible light region. Furthermore, it can be seen that there is no significant difference between the N-TiO2 and N-0.5%Fe-TiO2. Electron paramagnetic resonance (EPR) spectroscopy has been used to study the iron-doped titania system both for laboratory prepared samples and commercial pigments, which is a highly sensitive technique for examining paramagnetic species (levels of Fe <0.01% are detectable) and can give valuable information about the lattice site. The information obtained from EPR may provide detailed description both of the nature of the species and their co-ordination symmetries in the solid. It is well-known that transition-metal ions are generally paramagnetic due to their partially filled d orbits. At room temperature, Fe3+ can easily give the EPR signals, while Fe2+ can not, due to its very short relaxation time. For the undoped TiO2 and N-TiO2 (Figure 4 a,b), there are no signal of paramagnetic species. However, all samples after iron doping show the obvious signal at g ) 1.992, which can be attributed to Fe3+ substituted Ti4+ in the TiO2 lattice.6,8 Therefore, we conclude that the iron ion has been incorporated into the lattice of titania. There is some difference between the 0.5%Fe-TiO2 (Figure 4c) and N-0.5%Fe-TiO2 (Figure 4d). It can be observed that the signal of 0.5%Fe-TiO2 is broaden and the intensity is higher than that of N-0.5%Fe-TiO2. It is because that the surrounding environment of Fe3+ in the TiO2 lattice is different for the nitrogen and iron co-doped samples since there is nitrogen atoms existed around the Fe3+. However, when the iron doping concentration increases, the effect of nitrogen on the paramagnetic signal of Fe3+ is not evident, which is probably due to that the relative concentration of nitrogen around the

Nano-TiO2 Co-Doped with Nitrogen and Iron (III)

J. Phys. Chem. C, Vol. 111, No. 28, 2007 10621

Figure 4. EPR spectra of (a) TiO2, (b) N-TiO2, (c) 0.5%Fe-TiO2, (d) N-0.5%Fe-TiO2, (e) 1.0%Fe-TiO2, and (f) N-1.0%Fe-TiO2 measured at room temperature.

Figure 6. Photodecompositon activity of RB over xFe-TiO2 and N-xFe-TiO2 with different Fe3+-doping concentrations under visible light irradiation for 4 h.

Figure 5. The UV-diffuse reflectance spectra of xFe-TiO2 (a) and N-xFe-TiO2 (b) with different Fe3+-doping concentrations.

same iron for the N-1.0%Fe-TiO2 is decreased compared with the N-0.5%Fe-TiO2. UV-vis diffuse reflectance spectra of xFe-TiO2 and N-xFeTiO2 with different Fe3+ doping concentration are presented in panels a and b, respectively, of Figure 5. It is apparent that the diffuse reflectance spectra of all the samples doped with nitrogen and/or Fe3+ have extended a red shift and increased absorbance in the visible range. With the increasing of Fe3+-doping concentration, the absorbance increases. The red shift of the absorption edge in Fe3+-doped titania has been attributed to the charge-transfer transition between the iron ion d electrons and the TiO2 conduction or valence band.23 It has been reported that metal doping could form a dopant energy level within the band gap of TiO2. The electronic transitions from the valence

band to dopant level or from the dopant level to the conduction band can effectively red shift the band edge adsorption threshold.8 Furthermore, the absorption of samples co-doped with nitrogen and Fe3+ ion is abruptly stronger than that of the Fe3+-doped samples with the corresponding doping concentration of Fe3+. Since the starting absorption edge of co-doped sample shifts to the lower energy, the band gap of titania becomes narrowing compared to the undoped and doped samples with nitrogen or Fe3+ alone. 3.2. Photocatalytic Activities of Photocatalysts. The photocatalytic activity of rhodamine B (RB) was investigated by determining the remaining concentration of RB at various time intervals. Figure 6 shows the photodecomposition activity of RB over xFe-TiO2 and N-xFe-TiO2 with different Fe3+ doping concentration under visible light irradiation for 4 h after the dark adsorption for 30 min. Since the metal ion dopant can act as a mediator of interfacial charge transfer or as a recombination center, the efficiencies of the doped titania are changed. Many factors can affect the efficiency of photodecomposition, such as preparation method, doping concentration, energy level of the dopant within the TiO2 lattice, and the distribution of the dopant in the particle.9 Therefore, the doping concentration has a optimal value. In Figure 6, all the doped samples show the much higher photodecomposition activity compared with the pure TiO2. For the xFe-TiO2 photocatalysts, the photocatalytic activities increase with the Fe3+-doping concentration when the concentration is below 0.5 atom %. The optimal Fe3+-doping concentration is 0.5 atom % for the maximum photodecomposition of RB. Fe3+ doping becomes detrimental when the doping concentration is higher than the optimal value. The Fe3+ dopant can serve as a charge traps retarding the electron-hole combination rate and enhancing the interfacial charge transfer to degrade the RB within the suitable concentration range of dopant (lower than 0.5 atom %). When the dopant concentration is too high, the recombination rate will increase because the distance between trapping sites in a particle decreases. It is significant to point out that the photodegradation activities of TiO2 co-doped with nitrogen and Fe3+ ion are much higher than that of doped with the nitrogen or Fe3+ ion with the corresponding concentration alone. The probably mechanism is discussed in the following part. In Figure 7, the photodecomposition activities of RB over xFe-TiO2 and N-xFe-TiO2 with different Fe3+-doping concentration under UV light irradiation for 2 h are presented. When the Fe3+-doping concentration is less than 0.5%, the photocatalytic activities of the Fe3+-doped titania are slightly higher than

10622 J. Phys. Chem. C, Vol. 111, No. 28, 2007

Cong et al.

Figure 8. The schematic diagram of mechanism of the nitrogen and Fe3+ co-doping. Figure 7. Photodecomposition activity of RB over xFe-TiO2 and N-xFe-TiO2 with different Fe3+-doping concentration under UV light irradiation for 2 h.

that of the undoped titania, while that of the nitrogen and Fe3+co-doped titania are lower than that of the Fe3+-doped alone. When the Fe3+-doping concentration increases higher than 0.5%, the photocatalytic activities of Fe3+-doped titania decrease abruptly, while that of the codoped catalysts higher than that of the Fe3+-doped alone. 3.3.Probable Mechanism Discussion. Considering the effects of nitrogen doping and iron ions doping and the differences of photocatalytic activities between the nitrogen and/or Fe3+-doped titania, the probable schematic diagram is proposed. On the one hand, the co-doping with nitrogen and Fe3+ induces the band gap of TiO2 much narrowing. Anpo and co-workers have reported that in the metal ion-implanted TiO2 the overlap of the conduction band due to Ti (d orbital) of TiO2 and the metal (d orbital) orbital of the implanted metal ions could decrease the band gap of TiO2 to absorb the visible light.24 Furthermore, it is generally accepted that nitrogen doping can form a new states lie just above the valence band for the substitutional nitrogen or lie higher in the gap, which all could decrease the band gap of TiO2 and absorb the visible light. It can also be testified by the UV-vis diffuse spectra (Figure 5). Therefore, the nitrogen and Fe3+ ion doping induced the formation of new states closed to the valence band and conduction band, respectively. The co-operation of the nitrogen and Fe3+ ion leads to the much narrowing of the band gap and greatly improves the photocatalytic activity in the visible light region. On the other hand, the co-doping of nitrogen and Fe3+ ion inhibits the recombination of the photogenerated electron and hole. The suitable concentration range of Fe ion can trap the photogenerated electron, while the nitrogen can trap part of photogenerated holes, which enhances the utilization efficiency of photogenerated electron and hole. As illustrated in Figure 8, in the visible light irradiation, the electron cannot be excited from the valence band to the conduction band for the pure titania; i.e., the process A cannot occur. After nitrogen and Fe3+ doping, the electron can be excited from the N impurity level to the Fe3+ impurity level (process B) or from the N impurity level to the conduction band (process C) or from the valence band to the Fe3+ impurity level (process D). Therefore the quantity of the photoinduced electrons and holes is much higher than that of the pure titania and the nitrogen or Fe3+-doped titania. In a word, the probably mechanism of the improvement of the photocatalytic activity in visible light region by co-doping with nitrogen and Fe3+ ion is due to the band gap narrowing and the enhancement of the utilization efficiency of solar energy.

Figure 9. Photoluminescence emission spectra for (a) 0.5%Fe-TiO2 and (b) N-0.5%Fe-TiO2 powders measured at room temperature. Excitation wavelength ) 280 nm.

In the ultraviolet irradiation, the effects of doping on the process of photodecomposition may be different. We presume that there are electrons excited from the valence band to the conduction band (process A) for the pure titania under the ultraviolet irradiation. When there are nitrogen ions existed surrounding the Fe3+, the recombination possibility of electrons and holes is enhanced since the nitrogen and Fe3+ could trap holes and electrons, respectively. Therefore, although the codoping with nitrogen and iron ions narrows the band gap of the titania, the recombination of electrons and holes (process E) are relatively faster. It can be supported by the PL emission spectra, which have been widely used to investigate the efficiency of charge carrier trapping, migration, and transfer and to understand the fate of electron-hole pairs in semiconductor particles since PL emission results from the recombination of free carriers.25,26 Figure 9 shows the PL spectra of the 0.5%Fe-TiO2 (a) and N-0.5%Fe-TiO2 (b). It can be seen that the PL emission intensity of the nitrogen and 0.5% iron ions doped TiO2 is higher than that of the 0.5% iron ions doped TiO2. Because the PL emission is the result of the recombination of excited electrons and holes, the lower PL intensity of the doped sample indicates a lower recombination rate.24 Therefore, it can be deduced that the recombination efficiency of the nitrogen and 0.5% Fe3+ ions titania is relatively faster in the ultraviolet irradiation, which is probably the reason of the photocatalytic activity decreasing. In addition, another reason of the photocatalytic activities of nitrogen doping in the ultraviolet decrease may be the unsuitable of the nitrogen doping concentration, which is not the optimal value. Further experiments are needed. In a word, the recombination of the electrons and holes may be faster compared with the pure titania and become the controllable procedure for the

Nano-TiO2 Co-Doped with Nitrogen and Iron (III) nitrogen and iron ions co-doped titania in the ultraviolet photocatalytic reactions. The photocatalyst of nitrogen and 0.5% Fe3+-co-doped titania still shows the higher activity compared with the pure titania. 4. Conclusions In this study, photocatalysts TiO2 co-doped with nitrogen and Fe3+ are successfully synthesized by homogeneous precipitationhydrothermal process. The nitrogen doping can inhibit the growth and aggregation of the particles. Fe3+ can be incorporated into the lattice of titania and exist in states of Fe3+, which is testified by XRD, Raman and EPR spectra. There is little change in the phase structure after nitrogen and Fe3+ doping. The UV-diffuse reflectance spectra of the doped samples shift remarkably to the visible light region and the absorption enhance significantly. The photocatalytic activities of the nitrogen and/ or Fe3+-doped titania are investigated both under ultraviolet and visible light irradiation. The probable mechanism of co-doping effect is proposed. It is presumed that the co-operation of the nitrogen and Fe3+ ion leads to the much narrowing of the band gap and greatly improves the photocatalytic activity in the visible light region. Meanwhile, the co-doping can also promote the separation of the photogenerated electrons and holes to accelerate the transmission of photocurrent carrier. The photocatalyst co-doped nitrogen and 0.5% Fe3+ shows the best photocatalytic activity and the degradation efficiency were improved 75% and 5% under both visible and ultraviolet irradiation, respectively, compared with the pure titania. Acknowledgment. This work has been supported by Program for New Century Excellent Talents in University (NCET04-0414), Shanghai Nanotechnology Promotion Centre (0552nm019), National Nature Science Foundation of China (20577009), and the Ministry of Science and Technology of China (2006AA06Z379, 2006DFA52710). Part of the research was finished in Shanghai Nanotechnology Joint Lab. References and Notes
(1) Ranjit, K. T.; Willner, I.; Bossmann, S. H.; Braun, A. M. EnViron. Sci. Technol. 2001, 35, 1544.

J. Phys. Chem. C, Vol. 111, No. 28, 2007 10623


(2) Lin, Y.-M.; Tseng, Y.-H.; Huang, J.-H.; Chao, C. C.; Chen, C.-C.; Wang, I. Environ. Sci. Technol. 2006, 40, 1616. (3) Wang, X.; Yu, J. C.; Chen, Y.; Wu, L.; Fu, X. Environ. Sci. Technol. 2006, 40, 2369. (4) Zhang, Z;, Wang, C. -C; Zakaria, R; Ying, Y. J. Phys. Chem. B 1998, 102, 10871. (5) Zhu, J.; Deng, Z.; Chen, F.; Zhang, J.; Chen, H.; Anpo, M.; Huang, J.; Zhang, L. Appl. Catal., B 2006, 62, 329. (6) Zhu, J.; Zheng, W.; He, B.; Anpo, M. J. Mol. Catal. A 2004, 216, 35. (7) Di Paola, A.; Marci, G.; Palmisano, L.; Schiavello, M.; Uosaki, K.; Ikeda, S.; Ohtani, B. J. Phys. Chem. B 2002, 106, 637. (8) Nagaveni, K.; Hegde, M. S.; Madras, G. J. Phys. Chem. B 2004, 108, 20204. (9) Wonyong, C.; Andreas, T.; Michael, R. H. J. Phys. Chem. 1994, 98, 13669. (10) Yamashita, H.; Harada, M.; Misaka, J.; Takeuchi, M.; Ikeue, K.; Anpo, M. J. Photochem. Photobiol., A 2002, 148, 257. (11) Zhao, W.; Ma, W.; Chen, C.; Zhao, J.; Shuai, Z. J. Am. Chem. Soc. 2004, 126, 4782. (12) Sakthivel, S.; Kisch, H. Angew. Chem., Int. Ed. 2003, 42, 4908. (13) Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Science 2001, 293, 269. (14) Sato, S.; Nakamura, R.; Abe, S. Appl. Catal., A 2005, 284, 131. (15) Yu, J. C.; Ho, W.; Yu, J.; Yip, H.; Wong, P. K.; Zhao, J. EnViron. Sci. Technol. 2005, 39, 1175. (16) Yu, J. C.; Yu, J.; Ho, W.; et al. Chem. Mater. 2002, 14, 3808. (17) Umebayashi, T.; Yamaki, T.; Yamamoto, S.; Miyashita, A.; Tanaka, S.; Sumita, T.; Asai, K. J. Appl. Phys. 2003, 93, 5156. (18) Pottier, A.; Chaneac, C.; Tronc, E.; Mazerollers, L.; Jolivet, J-P. J. Mater. Chem. 2001, 11, 1116. (19) Berger, H.; Tang, H.; Kevy, F. J. Crys. Growth 1993, 130, 108. (20) Li Bassi, A.; Cattaneo, D.; Russo, V.; Bottani, C. E.; Bartoriri, E.; Mazza, T.; Pisen, P.; Midani, P.; Ernst, F. O.; Wegner, K.; Pratsinis, S. E. J. Appl. Phys. 2005, 98 (7), 074305. (21) Bersani, D.; Lottici, P. P.; Ding, X. Z. Appl. Phys. Lett. 1998, 72, 73-75. (22) Di Valentin, C.; Pacchioni, G.; Selloni, A.; et al. J. Phys. Chem. B 2005, 109, 11414. (23) Litter, M. I.; Navo, J. A. J. Photochem. Photobiol., A 1996, 98, 171. (24) Tang, H.; Prasad, K.; Sanjines, R.; Schmidd, P. E.; Levy, F. J. Appl. Phys. 1994, 75, 2042. (25) Li, F. B.; Li, X. Z. Chemosphere 2002, 48, 1103. (26) Zhang, J.; Hu, Y.; Matsuoka, M.; Yamashita, H.; Minagawa, M.; Hidaka, H.; Anpo, M. J. Phys. Chem. B 2001, 105, 8395.

Вам также может понравиться