Photophysical Properties and Photocatalytic Activities of Bismuth

Aug 19, 2006 - Photophysical Properties and Photocatalytic Activities of Bismuth Molybdates under Visible Light Irradiation ..... Especially paying at...
0 downloads 12 Views 865KB Size
17790

J. Phys. Chem. B 2006, 110, 17790-17797

Photophysical Properties and Photocatalytic Activities of Bismuth Molybdates under Visible Light Irradiation Yoshiki Shimodaira,† Hideki Kato,† Hisayoshi Kobayashi,‡ and Akihiko Kudo*,†,§ Department of Applied Chemistry, Faculty of Science, Science UniVersity of Tokyo, 1-3 Kagurazaka, Shinjyuku-ku, Tokyo 162-8601, Japan, Department of Chemistry and Materials Technology, Kyoto Institute of Technology, Matsugasaki, Sakyo-ku, Kyoto 606-8585, Japan, and Core Research for EVolutional Science and Technology, Japan Science and Technology Agency (CREST, JST), 4-1-8 Honcho, Kawaguchi-shi, Saitama 332-0012, Japan ReceiVed: April 11, 2006; In Final Form: July 21, 2006

Aurivillius structure Bi2MoO6 (BG: 2.70 eV) that is a low-temperature phase showed an intense absorption band in the visible light region and photocatalytic activity for O2 evolution from an aqueous silver nitrate solution under visible light irradiation, among various bismuth molybdates (Bi2MoO6, Bi2Mo2O9, and Bi2Mo3O12) synthesized by solid-state and reflux reactions. Bi2Mo3O12 (BG: 2.88 eV) also showed photocatalytic activity for O2 evolution under full-arc irradiation of a Xe lamp (λ > 300 nm). The photocatalytic activity of the Aurivillius structure Bi2MoO6 prepared by the reflux method was dependent on the annealing temperature after the preparation. The crystallinity was the important factor for the activity. Calculation by the density functional method indicated that the conduction band of Aurivillius structure Bi2MoO6 was made up of Mo 4d orbitals. It turned out that the visible-light absorption of this photocatalyst was due to the transition from the valence band consisting of O 2p orbitals to the conduction band. The corner-sharing structure of the MoO6 octahedra contributed to the visible light response and the photocatalytic performance because excitation energy and/or photogenerated electron and hole pairs began to migrate easily in the Aurivillius structure.

1. Introduction The photocatalytic water splitting into H2 and O2 which is an ideal way to get H2 has been an important reaction from the viewpoint of global energy and environmental issues.1-3 Recently, new photocatalysts consisting of metal ions with d0 and d10 electron configurations have been developed for photocatalytic H2 and O2 evolution from water under UV irradiation, one after another.1-5 Titanium of the fourth group and niobium and tantalum of the fifth group are the main components of d0 oxide photocatalysts.6,7 WO3 with the corner-sharing structure of WO6 octahedra is well-known to show high photocatalytic activity for O2 evolution from aqueous solutions containing electron acceptors under visible light irradiation as a representative photocatalyst of the sixth groups’ oxides.8-10 PbWO4 has recently been reported to be active for water splitting under UV irradiation.11 In contrast, although the oxides of molybdenum, which is from the same group as tungsten, have been used for many industrial reactions and reported as fixed catalysts for partial oxidation,12,13 there is no report that molybdenum is used as the main composition element of photocatalysts for water splitting into H2 and O2, except for some molybdates with a Scheelite structure which have activity for O2 evolution from an aqueous silver nitrate solution.14 Bismuth mixed oxides with the Aurivillius structure represented by (Bi2O2)2+(An-1BnO3n+1)2- (A ) Ba, Bi, Pb, etc., B * To whom correspondence should be addressed. E-mail: a-kudo@ rs.kagu.tus.ac.jp. Fax: +81-33235-2214. † Science University of Tokyo. ‡ Kyoto Institute of Technology. § Japan Science and Technology Agency.

Figure 1. Structures of (a) Aurivillius-type Bi2MoO6 and (b) WO3.

) Ti, Nb, W, Mo, etc.) possess unique layered structures in which perovskite slabs of (An-1BnO3n+1)2- are sandwiched between (Bi2O2)2+ layers as shown in Figure 1.15-20 The structure of the (An-1BnO3n+1)2- layer is similar to that of a WO3 photocatalyst because they have corner-shared octahedra structures. Dielectric,21-23 ion conductive,24,25 luminescent,26 and catalytic properties of these materials family have attracted attention.27 The authors have reported that some tungstates and titanates with Aurivillius structure are new photocatalysts for H2 or O2 evolution.28 Moreover, BiVO4 was found to be a photocatalyst which is active for O2 evolution from an aqueous silver nitrate solution under visible light irradiation.29 Some bismuth oxides such as Bi2WO6 have recently been reported as photocatalysts for degradation of organic pollutants.30 Therefore, the photocatalytic properties of the bismuth mixed oxides with Aurivillius structure and the related compounds are also of interest.

10.1021/jp0622482 CCC: $33.50 © 2006 American Chemical Society Published on Web 08/19/2006

Bismuth Molybdate Photocatalysts In this paper, photophysical and photocatalytic properties of oxides consisting of a Bi3+ ion with 6s2 configuration and Mo6+ ion with 4d0 configration were studied to develop photocatalysts with visible-light response. The effects of the synthetic conditions and phase transition on the photocatalytic activities were examined. The energy structures were also discussed by the plane-wave-based density functional method calculation. 2. Experimental Section Bi2MoO6 was synthesized by a reflux method and a solidstate reaction.31 Bi2Mo2O9 and Bi2Mo3O12 were synthesized only by solid-state reactions.32 The starting materials used for the reflux reaction were Bi(NO3)3‚5H2O (Kanto Chemical; purity >99.5%) and H2MoO4 (Kanto Chemical). The starting materials used for the solid-state reaction were Bi2O3 (Kanto Chemical; purity >99.9%) and MoO3 (Kanto Chemical; purity >99.5%). In the case of a solid-state reaction, the starting materials were mixed on an alumina mortar and the mixtures were calcined at 673-973 K for 5 h in air using an alumina crucible (purity: 99.7%). In the case of a reflux reaction, the starting materials were soaked in water in a stoichiometric ratio, refluxed on a mantle heater for 24 h. The products were washed with water and filtered. The obtained precursors were dried at 320 K for 15 h in an oven and calcined at several temperatures (Bi2MoO6, 573-973 K for 5 h; Bi2Mo2O9, 823 K for 5 h; Bi2Mo3O12, 773973 K for 5 h) in air using an alumina crucible. We have especially paid attention to the phase transition temperature for the calcination of Bi2MoO6. The purity of the obtained powders were confirmed by X-ray diffraction (Rigaku: RINT1400, MiniFlex). Photocatalytic O2 evolution from an aqueous silver nitrate solution and H2 evolution from an aqueous methanol solution were carried out in a gas-closed circulation system. Catalyst powder (0.3 g) was dispersed by a magnetic stirrer in a reactant solution (0.05 mol/L AgNO3(aq) or 10 vol % MeOH(aq), 150 mL) in a cell with a side-window made of Pyrex. The light source was a 300-W Xe lamp (ILC technology; Cermax LX300F). Cutoff filters (Hoya) were employed for controlling the wavelength of incident light. Amounts of evolved gases were determined using gas chromatography (Shimazu, GC-8A, MS5A, TCD, Ar carrier). Diffuse reflection spectra were obtained using a UV-vis-NIR spectrometer (Jasco; UbestV-570) and were converted from reflection to absorbance by the KubelkaMunk method. Surface areas were determined by BET measurements (Coulter; SA3100). Photocatalyst powders were observed by a scanning electron microscope (Hitachi; S-5000). Thermogravimetric analysis and phase transition were examined by a thermogravimetry differential thermal analyzer (Ulvac; TGD9600). The plane-wave-based density functional method calculation was carried out for the low-temperature phase Bi2MoO6 with Aurivillius structure (AS-Bi2MoO6), high-temperature phase Bi2MoO6 (HTP-Bi2MoO6), Bi2Mo2O9, and Bi2Mo3O12 by employing a CASTEP program.33 The core electrons were replaced with ultrasoft core potentials,34 and the valence electronic configurations for Bi, Mo, and O atoms were 6s26p3, 4s24p65s14d5, and 2s22p4, respectively. The calculations were carried out using the conventional unit cells of [Bi2MoO6]2 for ASBi2MoO6, [Bi2MoO6]16 for HTP-Bi2MoO6, [Bi2Mo2O9]16, and [Bi2Mo3O12]4. The total numbers of the electrons were 120, 960, 736, and 496, and the numbers of the occupied orbitals were 60, 480, 368, and 248, respectively. 3. Results and Discussion 3.1. Diffuse Reflectance Spectrum. Diffuse reflectance spectra of bismuth molybdates are shown in Figure 2. The band

J. Phys. Chem. B, Vol. 110, No. 36, 2006 17791

Figure 2. Diffuse reflectance spectra of (a) AS-Bi2MoO6, (b) HTPBi2MoO6, (c) Bi2Mo2O9, and (d) Bi2Mo3O12.

gaps of AS-Bi2MoO6, HTP-Bi2MoO6, Bi2Mo2O9, and Bi2Mo3O12 were estimated to be 2.70, 3.02, 3.10, and 2.88 eV from the onsets of absorption edges, respectively. AS-Bi2MoO6 of a low-temperature phase with a corner-sharing structure of MoO6 octahedra18 possessed the onset of the absorption edge at the longest wavelength among them. HTP-Bi2MoO6 and Bi2Mo2O9 with isolated MoO4 tetrahedra18,32 showed the onsets at the shortest wavelength. Bi2Mo3O12 with both MoO4 tetrahedra and MoO6 octahedra32 showed just the onset of the intermediate region. The less the coordination number of oxygen ions to a molybdenum ion was, the wider the band gap was. This is explained by the interaction of oxygen and molybdenum orbitals. The less the coordination number is, the stronger the interaction between oxygen ions and a molybdenum ion is. The strong interaction makes the energy gap between a conduction band consisting of Mo 4d and a valence band consisting of O 2p wide. This results in the band gaps of the samples consisting of MoO4 tetrahedra being wider than those consisting of MoO6 octahedra. Moreover, the widths of the conduction bands of the samples consisting of MoO4 tetrahedra would be narrower than those consisting of MoO6 octahedra because the d-d splitting affecting the conduction bandwidth in MoO4 tetrahedra is smaller than that in MoO6 octahedra judging from crystal field theory. This effect also makes the band gap wider if the valence band positions are similar to each other. Therefore, we concluded that the band gap was mainly influenced by the coordination number of oxygen ions to a molybdenum ion. 3.2. Energy Structure. The band structures of low-temperature phase AS-Bi2MoO6, HTP-Bi2MoO6, Bi2Mo2O9, and Bi2Mo3O12 were studied by the plane-wave-based density functional method to clarify which orbitals contribute to valence band and conduction band formation. Figure 3a,b shows the band structure, the density of states (DOS), and the density contour maps for the LUMO and HOMO of low-temperature phase AS-Bi2MoO6. It was estimated that the valence band (#A25-#A60) was mainly derived from the O 2p orbitals and the conduction band (#A61-#A70) was derived from the Mo 4d orbitals and the Bi 6p orbitals. The contribution of the Bi 6s orbitals to the valence band formation, as seen in BiVO4 and Bi2WO6, was not observed.29,30 The O 2p orbitals along the direction of the layer of MoO6 octahedra mainly contributed to the formation of the top of the valence band (HOMO) rather than that in the vertical direction. Moreover, the contribution of O 2p orbitals of the Bi2O2 layer was not observed for the HOMO. The Bi 6p orbitals for the LUMO were not detected when the isosurface in the density of states

17792 J. Phys. Chem. B, Vol. 110, No. 36, 2006

Shimodaira et al.

Figure 3. (a) Band structure and total density of state and (b) density contour maps for the LUMO and HOMO for AS-Bi2MoO6 calculated by the density functional method.

was 0.3. They were observed when the isosurface in the density of states was 0.1. Hence, the contribution of each orbital was not equivalent, although the LUMO consisted of hybrid orbitals of Mo 4d and Bi 6p orbitals. The Mo 4d orbitals mainly contributed to the LUMO rather than the Bi 6p orbital. The main orbitals consisting of the conduction band were Mo 4d orbitals. Visible light absorption of AS-Bi2MoO6 was revealed to be due to the transition from the valence band consisting of the O 2p orbitals to the conduction band derived from the primary Mo 4d orbitals in MoO6 octahedra and the secondary Bi 6p orbitals. All valence bands of HTP-Bi2MoO6, Bi2Mo2O9, and Bi2Mo3O12 mainly consisted O 2p orbitals as well as occurred for AS-Bi2MoO6 as shown in Figures 4-6. However, the tops of the valence bands including the HOMOs of HTP-Bi2MoO6 (#B477-#B480), Bi2Mo2O9 (#C366-#C368), and Bi2Mo3O12 (#D247-#D248) consisted not only of the O 2p orbitals but also of Bi 6s orbitals partly. All conduction bands of the three materials consisted of Mo 4d orbitals. In contrast to AS-Bi2MoO6, Bi 6s

orbitals contributed to raising their valence bands resulting in a narrowing of their band gaps. Especially paying attention to the Mo 4d orbitals forming the conduction band (#D249-#D270) of Bi2Mo3O12, which contains both MoO4 tetrahedra and MoO6 octahedra, by analyzing density contour maps, one finds the Mo 4d orbitals of MoO6 octahedra widely contributed to the whole conduction band (#D249-#D270), whereas the MoO4 tetrahedra contributed to only the middle of the conduction band (#D257-#D263). This means that the Mo 4d orbitals of the MoO6 octahedra generate wide conduction bands. It is also one of the reasons that the band gap of a material containing the MoO6 octahedra was smaller than that containing the MoO4 tetrahedra. The calculated band gaps of AS-Bi2MoO6, HTP-Bi2MoO6, Bi2Mo2O9, and Bi2Mo3O12 were 0.68, 2.86, 3.07, and 2.42 eV, respectively. The order in the calculated band gaps corresponded to the values experimentally determined from the diffuse reflectance spectrum. Distribution of the band structure gives information on the mobility of excitation energy.35 The distribution of AS-Bi2-

Bismuth Molybdate Photocatalysts

J. Phys. Chem. B, Vol. 110, No. 36, 2006 17793

Figure 4. (a) Band structure and total density of state and (b) density contour maps for the LUMO and HOMO for HTP-Bi2MoO6 calculated by the density functional method.

MoO6 was larger than those of HTP-Bi2MoO6, Bi2Mo2O9, and Bi2Mo3O12. This property is advantageous for photocatalytic performance. 3.3. Photocatalytic Activities. The photocatalytic activities of bismuth molybdates are summarized in Table 1. Lowtemperature phase AS-Bi2MoO6 and Bi2Mo3O12 showed reasonable activities for O2 evolution from an aqueous silver nitrate solution under visible light irradiation. If this is expressed in another way, only these materials containing MoO6 octahedra in the structure showed the photocatalytic activity whereas those containing MoO4 tetrahedra showed negligible photocatalytic activities among the bismuth molybdates tested.

It is considered that the reason AS-Bi2MoO6 showed higher activity than Bi2Mo3O12 is the crystal structure. AS-Bi2MoO6 possesses perovskite-like slabs consisting of MoO6 octahedra of which the structure is often advantageous for photocatalysts36 because excitation energy can easily migrate in the structure as suggested from the distribution of the band structure as shown in Figure 2a. Another reason was the difference in the number of absorbed photon because of the different band gaps. On the other hand, Bi2Mo3O12 has a certain connection between MoO4 tetrahedra and MoO6 octahedra while the degree of isolation of MoO4 tetrahedra is large in HTP-Bi2MoO6 and Bi2Mo2O9. Therefore, the degree of the localization of photogenerated

17794 J. Phys. Chem. B, Vol. 110, No. 36, 2006

Shimodaira et al.

Figure 5. (a) Band structure and total density of state and (b) density contour maps for the LUMO and HOMO for Bi2Mo2O9 calculated by the density functional method.

electron-hole pairs in HTP-Bi2MoO6 and Bi2Mo2O9 is larger than that in Bi2Mo3O12. It results in that photocatalytic activities of HTP-Bi2MoO6 and Bi2Mo2O9 were negligible. No bismuth molybdates were active for H2 evolution from an aqueous methanol solution. From the band calculation, the bottom-up effect of the valence band by Bi 6s orbitals as seen in the BiVO4 photocatalyst material29 was not observed for ASBi2MoO6, indicating that the top of the valence band was formed by only O 2p orbitals. Therefore, the conduction band level of AS-Bi2MoO6 does not seem to be enough for water reduction to form H2, judging from the valence band character and the band gap of a d0 metal oxide.37 We believe that the development of new photocatalysts with activities even for a half reaction of water splitting, H2 or O2 evolution in the presence of sacrificial regents, is important. This material actually contributes to the

construction of a Z-scheme for overall water splitting under visible light irradiation.38 Figure 7 shows the X-ray diffraction patterns of Bi2MoO6 powders obtained by calcining the precursor prepared by a reflux method. A single low-temperature phase AS-Bi2MoO6 was obtained even as prepared. The Aurivillius phase was observed by calcination at wide range temperatures (373-823 K). The phase transition from the Aurivillius phase to the hightemperature phase, HTP-Bi2MoO6, was observed at 873 K. The single high-temperature form was obtained above 923 K. This result agreed with the TG-DTA measurement reported by Kodama et al.39 In the phase transition of Bi2MoO6, there is an irreversible second-order phase transition from ferroelectrics to paraelectrics with anomalous specific heat thereby transforming the layer structure containing MoO6 octahedra into a bulk

Bismuth Molybdate Photocatalysts

J. Phys. Chem. B, Vol. 110, No. 36, 2006 17795

Figure 6. (a) Band structure and total density of state and (b) density contour maps for the LUMO and HOMO for Bi2Mo3O12 calculated by the density functional method.

TABLE 1: Photocatalytic O2 Evolution over Bismuth Molybdates photocatalysta

band gap/eV

AS-Bi2MoO6 reflex method HTP-Bi2MoO6 Bi2Mo2O9 Bi2Mo3O12

2.70 3.02 3.10 2.88

incident light/ nm

activityb O2/µmol h-1

λ > 420 λ > 300 λ > 300 λ > 300 λ > 420 λ > 300

55 127 0.7 1.8 7.6 46

a 0.5 g. b Reactant solution: an aqueous silver nitrate solution (0.05 mol L-1), 150 mL. Light source: 300 W Xe lamp.

structure containing MoO4 tetrahedra.18,40,41 The half-width of a peak at 28° was decreased as the calcination temperature was higher than 773 K as shown in Table 2. This indicated that the crystallinity became good. At the same time, the surface area

was drastically decreased from 41.1 to 0.9 m2/g. Figure 8 shows SEM images of the low-temperature phase (AS-Bi2MoO6) as prepared and calcined at 823 K. The particle shape of noncalcined Bi2MoO6 was platelike, and the size was 20-100 nm. The particle size of the sample calcined at 823 K was 1-3 µm by sintering. The sintered particles consisted of aggregates of primary particles. Shifts of absorption edges of diffuse reflectance spectra were observed by the calcination at wide temperature ranges as shown in Figure 9. The absorption edge was shifted to the longwavelength side and became steep as the calcination temperature was increased from 373 to 823 K. The change in the DRS is due to the crystallinity as observed by XRD. The narrow band gap was obtained for AS-Bi2MoO6 calcined at 823 K, which possessed high crystallinity. AS-Bi2MoO6 showed the highest photocatalytic activity. On the other hand, the absorption edge

17796 J. Phys. Chem. B, Vol. 110, No. 36, 2006

Shimodaira et al.

Figure 7. X-ray diffraction patterns of Bi2MoO6 calcinated at various temperatures: (a) 373 K; (b) 673 K; (c) 773 K; (d) 803 K; (e) 823 K; (f) 853 K; (g) 863 K; (h) 873 K; (i) 923 K; (j) 973 K.

TABLE 2: Surface Area and Half-Width of a Peak at 28° of Bi2MoO6 Powder Calcined at Different Temperatures calcination temp/K

surf area/ m2 g-1

half-width of a peak at 28°/deg

373 (uncalcination) 573 673 773 803 823

41.1 1.2 1.1 0.9

0.304 0.259 0.246 0.188 0.169 0.162

0.9

was blue-shifted by the phase transition around 823 K of the calcination temperature. The absorption edge was at the shortest wavelength when the single phase of HTP-Bi2MoO6 was obtained. Figure 10 shows the relationship between calcination temperature and photocatalytic oxygen evolution of Bi2MoO6 prepared by a reflux method. Bi2MoO6 calcined at 823 K showed the highest photocatalytic activity for O2 evolution under visible light irradiation. Moreover, the decrease in the photocatalytic activity corresponded to an irreversible phase transition as shown in Figure 7. The improvement of the photocatalytic activity by calcination around 823 K was due to an increase in crystallinity of Aurivillius structure as shown in Table 2. The half-width of the peak at 28° for the sample calcined at 803 K was quite narrow compared with that at 773 K. Moreover, the former was similar to that at 823 K. This indicated that an increase in the crystallinity was almost completed at 803 K although the crystallinity of the sample calcined at 823 K was slightly better than that at 803 K. This is the reason the photocatalytic activity suddenly increased around 803 K of the calcination. This result indicated that the crystallinity was the important factor rather than the surface area for this photocatalytic reaction.

Figure 8. Scanning electron microscope images of Bi2MoO6. Acceleration voltage: 10 kV. Key: (a) only refluxed at 373 K; (b) refluxed at 373 K and calcined at 823 K.

Figure 9. Diffuse reflectance spectra of Bi2MoO6 prepared by the reflux method calcined at various temperatures: (a) 373 K; (b) 823 K; (c) 853 K; (d) 923 K.

There was no difference in the photocatalytic activity for the O2 evolution between catalysts prepared by a solid-state reaction and a reflux method. AS-Bi2MoO6 was also prepared by only calcination at higher than 773-823 K in the solid-state reaction. In contrast to the solid-state reaction, it was able to be prepared by calcinations between 373 and 823 K in the reflux method. It is noteworthy that the reflux method made us investigate the

Bismuth Molybdate Photocatalysts

J. Phys. Chem. B, Vol. 110, No. 36, 2006 17797 the photocatalytic ability of AS-Bi2MoO6. The reflux method made the investigation of the photocatalytic activity of AS-Bi2MoO6 prepared at a wide range of calcination temperature possible. This investigation revealed the importance of crystallinity for the photocatalytic performance of AS-Bi2MoO6. Acknowledgment. This work has been supported by the Core Research for Evolutional Science and Technology (CREST) of the Japan Science and Technology Agency (JST) and a Grantin-Aid (No. 14050090) for the priority Area Research (No. 417) from the Ministry of Education, Culture, and Technology. References and Notes

Figure 10. Photocatalytic oxygen evolution under visible light irradiation over Bi2MoO6 calcinated at several temperatures: (a) lowtemperature phase, Aurivillius structure; (b) high-temperature phase. Sample (0.5 g): an aqueous silver nitrate solution (0.05 mol L-1), 150 mL. Light source: 300-W Xe lamp with a cutoff filter (λ > 420 nm). Cell: side irradiation type reaction vessel made of Pyrex.

Figure 11. Diffuse reflectance spectrum and dependence of photocatalytic O2 evolution from an aqueous silver nitrate solution over ASBi2MoO6 photocatalyst prepared by the reflex method upon irradiated wavelength. Sample (0.5 g): an aqueous silver nitrate solution (0.05 mol L-1), 150 mL. Light source: 300-W Xe lamp with cutoff filters. Cell: side irradiation type reaction cell made of Pyrex.

relationship between the photocatalytic properties of Bi2MoO6 and the calcination temperature at a wide range. Figure 11 shows the diffuse reflectance spectrum and the dependence of O2 evolution from an aqueous silver nitrate solution over the AS-Bi2MoO6 photocatalyst prepared at optimal condition upon cutoff wavelength using cutoff filters. The horizontal axis indicates the cutoff wavelength in which transmittance is almost 0%. The onset of the wavelength dependency of the photocatalytic activity agreed with that of the diffuse reflectance spectrum indicating that the O2 evolution reaction photocatalytically proceeded over Bi2MoO6 by band gap excitation. 4. Conclusions AS-Bi2MoO6 (BG: 2.70 eV) consisting of a layered structure with corner-shared MoO6 units has arisen as a new visible lightdriven photocatalyst for O2 evolution from an aqueous silver nitrate solution. This study succeeded in the discovery of molybdenum oxide photocatalysts for O2 evolution from aqueous solutions under visible light irradiation. AS-Bi2MoO6 possesses the corner-sharing structure of MoO6 octahedra as seen in the well-known WO3 photocatalyst with visible light response. This structure affected the energy gap between HOMO and LUMO, the width of the conduction band, and the delocalization of excitation energy, which contributed to the band structure and

(1) Kudo, A.; Kato, H.; Tsuji, I. Chem. Lett. 2004, 33, 1534. (2) Domen, K.; Kondo, N, J.; Hara, M.; Takata, T. Bull. Chem. Soc. Jpn. 2000, 73, 1307. (3) Sato, J.; Kobayashi, H.; Ikarashi, K.; Saito, N.; Nishiyama, H.; Inoue, Y. J. Phys. Chem. B 2004, 108, 4369. (4) Ikarashi, K.; Sato, J.; Saito, N.; Nishiyama H.; Inoue, Y. J. Phys. Chem. B 2002, 106, 9048. (5) Sato, J.; Saito, N.; Nishiyama H.; Inoue, Y. J. Phys. Chem. B 2003, 107, 7965. (6) Kato, H.; Kudo, A. J. Phys. Chem. B 2002, 106, 5029. (7) Kato, H.; Kobayashi, H.; Kudo, A. J. Phys. Chem. B 2002, 106, 12441. (8) Krasnovsky, A. A.; Brin, G. P. Dokl. Akad. Nauk. 1962, 147, 656. (9) Darwent, J. R.; Mills, A. J. Chem. Soc., Faraday Trans. 1982, 2, 359. (10) Erbs, W.; Desilvestro, J.; Borgarello, E.; Gratzel, M. J. Phys. Chem. 1984, 88, 4001. (11) Saito, N.; Kadowaki, H.; Kobayashi, H.; Ikarashi, K.; Nishiyama, H.; Inoue, Y. Chem. Lett. 2004, 33, 1452. (12) Batist, P. A.; Der kinderen, A. H. W. M.; Leeuwenburgh, Y.; Metz, F. A. M. G.; Schuit, G. C. A. J. Catal. 1968, 12, 45. (13) Batist, P. A.; Bouwens, J. F. H.; Schuit, G. C. A. J. Catal. 1972, 25, 1. (14) Kato, H.; Matsudo, N.; Kudo, A. Chem. Lett. 2004, 33, 1216. (15) Newnmam, R. E.; Wolfe, R. W.; Dorrian, J. F. Mater. Res. Bull. 1971, 6, 1029. (16) Hutchison, J. L.; Anderson, J. S.; Rao, C. N. R. Proc. R. Soc. London, Ser. A 1977, 355, 301. (17) Bando, Y.; Watanabe, A,; Sekikawa, Y.; Goto, M.; Horiuchi, S. Acta Crystallogr., Sect. A 1979, 35, 142. (18) Frit, B.; Mercorio, J. P. J. Alloys Compd. 1992, 188, 27. (19) Montero, M. T.; Millan, P.; Duran-Martin, P.; Jimenez, B. Catro, A. Mater. Res. Bull. 1998, 33, 1103. (20) Watanabe, A.; Goto, M. J. Less-Common Met. 1978, 61, 265. (21) Cummins, S. E.; Cross, L. E. J. Appl. Phys. 1968, 39, 2268. (22) Hamada, M.; Tabata, H.; Kawai, T. Thin Solid Films 1997, 306, 6. (23) Hirose, T.; Kawaminami, M.; Obara, K.; Arakawa, M.; Takeuchi, H. J. Phys. Soc. Jpn. 1993, 62, 1758. (24) Baux, N.; Vannier, R. N.; Mairesse, G.; Nowogrocki, G. Solid State Ionics 1996, 91, 243. (25) Islam, M. S.; Lazure, S.; Vannier, R. N.; Nowogrocki, G.; Mairesse, G. J. Mater. Chem. 1998, 8, 655. (26) Bordun, O. M. Inorg. Mater. 1998, 34, 1270. (27) Sinhamahapatra, P. K.; Sinhamahapatra, S.; Bhattacharyya, S. K. Krist. Tech. 1977, 12, 1205. (28) Kudo, A.; Hijii, S. Chem. Lett. 1999, 1103. (29) Kudo, A.; Omori, K.; Kato, H. J. Am. Chem. Soc. 1999, 121, 11459. (30) Fu, H.; Pan, C.; Yao, W.; Zhu, Y. J. Phys. Chem. B 2005, 109, 22432. (31) Kodama, H.; Izumi, F.; Watanabe, A. J. Solid State Chem. 1981, 36, 349. (32) Chen, T. J. Cryst. Growth 1973, 20, 29. (33) Payne, M. C.; Teter, M. P.; Allan, D. C.; Arias, T. A.; Joannopoulos, J. D. ReV. Mod. Phys. 1992, 64, 1045. (34) Vanderbilt, D. Phys. ReV. 1990, B41, 7892. (35) Sato, J.; Kobayashi, H.; Inoue, Y. J. Phys. Chem. B 2003, 107, 7970. (36) Kato, H.; Kudo, A. J. Phys. Chem. B 2001, 105, 4285. (37) Scaife, E. D. Sol. Energy 1980, 25, 41. (38) Kato, H.; Hori, M.; Konta, R.; Shimodaira, Y.; Kudo, A. Chem. Lett. 2004, 33, 1348. (39) Kodama, H.; Watanabe, A. J. Solid State Chem. 1985, 56, 225. (40) Buttrey, D. J.; Vogt, T.; White, B. D. J. Solid State Chem. 2000, 155, 206. (41) Ramaswamy, M.; Raje, G.; Jayaraman, K.; Shajina, S.; Balan, P.; Sriramulu, M. J. Phys.: Condens. Matter 2002, 14, 4001.