Improving photocatalytic properties of SrTiO3 through (Sb, N) codoping: a hybrid density functional study

Brindaban Modak, K. Srinivasu and Swapan K. Ghosh*
Theoretical Chemistry Section, Bhabha Atomic Research Centre, Homi Bhabha National Institute, Mumbai – 400 085, India. E-mail: skghosh@barc.gov.in; Tel: +91-22-25595092

Received 18th July 2014 , Accepted 10th September 2014

First published on 10th September 2014


Abstract

A systematic study using hybrid density functional theory has been carried out to investigate the synergistic effect of Sb and N doping on the photocatalytic properties of SrTiO3 under visible light. The calculated band gap (3.19 eV) for SrTiO3 with the Heyd, Scuseria, and Ernzerhof hybrid functional is found to be very close to the experimentally observed value of 3.2 eV. Although doping with N is able to enhance the visible light activity by reducing the effective band gap to 2.31 eV, the localized occupied and unoccupied states in the forbidden region may affect the photocatalytic activity. However, the presence of Sb not only passivates those unoccupied states completely, but also shifts the localized occupied states near the valence band to form a continuum band structure. The introduction of N into the SrTiO3 crystal structure is favored by the presence of Sb. In the codoped system charge compensation is established, thereby unwanted vacancy formation will be minimized. The absorption curve for the (Sb, N)-codoped SrTiO3 is found to shift towards the visible region due to reduction in band gap to 2.66 eV. Moreover, the band alignment shows that the (Sb, N)-codoping makes SrTiO3 thermodynamically more suitable for hydrogen production as compared to the undoped system. Based on the present study, we can propose (Sb, N)-codoping as one of the effective approaches to improve the photocatalytic activity of SrTiO3 for water splitting under visible light irradiation.


1. Introduction

Photo splitting of water under sunlight is one of the most promising ways to generate hydrogen which has been widely accepted as an alternative energy carrier.1 The biggest challenge is to find a suitable catalyst for this purpose. To date numerous efforts have been made towards developing semiconductor based photocatalysts.2–4 Among the perovskites, SrTiO3 has been extensively studied for its potential applications in this field. Since the band gap of SrTiO3 is 3.2 eV, its photoactivity is limited to the UV region of the solar spectrum.5 There have been made several experimental as well as theoretical attempts by doping with foreign element/elements to improve the visible light activity of SrTiO3. These include cationic doping either at the Sr lattice site6 or at the Ti lattice site7–16 as well as anionic doping at the oxygen lattice site.17–23 Among the transition metal doped systems, Cr-doped SrTiO3 has been extensively studied due to its potential application to split water and organic compounds under visible light.7–11 However, there exists some major issues that limit the photocatalytic efficiency of Cr-doped SrTiO3. Improved photocatalytic activity has been observed only when chromium is in trivalent state (Cr3+), although chromium is more stable in hexavalent state (Cr6+).8 Hence keeping chromium in trivalent state is a challenge as it has tendency to be converted in the hexavalent state. Secondly, localized mid gap states introduced due to Cr doping not only hinder the mobility of the charge carriers, but also bring significant changes in the band edge positions, thus affect overall water splitting property.11 Owing to attractive visible light activity Rh-doped SrTiO3 has been found to be studied by several groups.12–16 However, photoactivity of Rh-doped SrTiO3 is limited by the presence of Rh4+ species, which plays significant role in the electron–hole recombination.16 Doping with non-metal elements are also extensively studied for the band gap engineering of the oxide based semiconductor photocatalysts. Among them N-doping has attracted immense interest for extending the absorption curve to the desirable range. Wang et al. observed that the photocatalytic activity under visible light is significantly improved due to N doping and the visible light absorption increases with increasing amount of N into the SrTiO3 crystal structure.21 However, this is due to the presence of localized N 2p states which appear above the valence band (VB) reducing the effective band gap.22 Another major challenge in case of monodoping is to avoid charge compensating defects, which are well known to diminish photocatalytic efficiency. There have been shown many efforts using codoping approach to overcome the undesirable consequences associated with those localized states and the additional charge introduced due to N-doping. Miyauchi et al.24 and Wang et al.25 successfully synthesized (La, N)-codoped SrTiO3, which shows improved photocatalytic activity under visible light. However, this photoactivity has been found to be limited to the decomposition of organic compounds. Although, it has been shown in the theoretical study of Wei et al. that codoping of either nonmetal (H, F, Cl, Br, I) or metal (V, Nb, Ta, Sc, Y, La) is able to passivates the N-induced discrete states, their applicability for the photo-splitting of water is still to be explored.26 Recently, Yu et al. synthesized (Cr, N)-codoped SrTiO3, which has been shown to generate only H2 during water splitting under visible light.27 In the present study, we employ Sb as codopant to improve the photoactivity of N-doped SrTiO3 under visible light. The advantages for choosing Sb as codopant are: (a) being very similarity in ionic radius (Sb5+: Ref = 0.60 Å; Ti4+: Ref = 0.605 Å),28 Sb can be easily fitted at the Ti lattice site without causing major lattice distortion in the host crystal structure; (b) introduction of Sb into the N-doped system is expected to maintain charge balance in the codoped SrTiO3 (Ti4+ + O2− = Sb5+ + N3−); (c) use of a non transition metal (Sb) may be more preferable over transition metal to avoid localized ‘d’-states in the forbidden region. It has been successfully shown in the earlier reports that introduction of Sb as a codopant significantly increases the photocatalytic property of the monodoped entity.16,29–36 As for example, codoping of Sb into the Rh-doped SrTiO3 effectively reduces the electron–hole recombination rate by fixing Rh oxidation state to +3 forming a charge compensated system.16 Similarly, codoping of Sb into the Cr-doped SrTiO3 has been found to enhance the photoactivity for H2 evolution by suppressing the formation of hexavalent chromium (Cr6+) as well as vacancy, which are known to diminish the photoactivity of SrTiO3.29 However, detailed theoretical studies exploring the effect of Sb on the electronic structure of SrTiO3 are rare. Here, we present a systematic study to investigate the synergistic effect of both N (as a substituent of O) and Sb (as a substituent of Ti) as codopants, and compare the results with that of the N-doped, Sb-doped and undoped SrTiO3, using density functional theory as a tool. Since the use of Heyd, Scuseria, and Ernzerhof (HSE)37 functional is known to successfully reproduce the experimental band gap for undoped SrTiO3, we employ the same functional in our calculations. Conclusions have been drawn by analyzing the band structure, partial density of states (PDOS), optical spectrum, and band alignment with respect to the water redox levels.

2. Computational methods

Spin-polarized DFT calculations have been carried out using Vienna ab initio simulation package (VASP)38 electronic structure code. The electron-ion interaction has been treated by projector augmented wave (PAW) potential,39 which is widely used for the studies of electronic structure of semiconductor materials. For the Brillouin zone integration we employ Monkhorst and Pack scheme to generate gamma-centered k-point sets.40 The mono-doped system is modeled by replacing either one of the Ti or O atoms by Sb or N, respectively from a 2 × 2 × 2 supercell (40 atoms) of the cubic (space group: Pm[3 with combining macron]m) SrTiO3 crystal structure. In the case of codoping, we simultaneously introduce both Sb and N into SrTiO3. This corresponds to Sb concentration of 12.5% and N concentration of 4.17%. The exchange and correlation energy density functionals were defined by generalized gradient approximation (GGA) with Perdew–Burke–Ernzerhof (PBE) scheme41 during geometry optimization. K-point mesh of 8 × 8 × 8 was found to be sufficient to attain convergence. The total energy convergence tolerance was set to 10−6 eV per atom for the self consistent iteration. Electronic structure calculations have been carried out by employing the HSE hybrid functional, where the short-range interactions are calculated with exact exchange mixing in both HF and DFT. The exchange-correlation energy is expressed as
 
EHSEXC = aESRX(μ) + (1 − a)EPBE,SRX(μ) + EPBE,LRX(μ) + EPBEC (1)
where, ‘a’ stands for the mixing coefficient. The screening parameter μ defines the short ranged (SR) and long ranged (LR) part of the interaction. In the present study, we use the mixing exchange parameter of 28% and standard screening parameter of 0.2 Å−1 to reproduce the experimental band gap of SrTiO3.11 K-point mesh of 3 × 3 × 3 was set for the hybrid functional calculations. The valence states considered during the calculations are: Sr (4s24p65s2), Ti (4s23d2), Sb (5s25p3), O (2s22p4), and N (2s22p3). The energy cutoff of 600 eV has been chosen for the plane wave basis sets. To investigate the absorption behavior frequency-dependent dielectric function calculation has been carried out for the codoped and undoped SrTiO3.

3. Results and discussion

Salient features of the electronic structure of undoped SrTiO3, monodoped SrTiO3 using N and Sb as dopants, and (Sb, N)-codoped SrTiO3 are discussed below.

3.1. SrTiO3

Cubic SrTiO3 is described by a framework of TiO6 perfect octahedron, where the center position is occupied by Ti. Fig. 1a shows the band structure plot for the undoped SrTiO3 along high symmetric k-path of the Brillouin zone (the horizontal dashed line represents the Fermi level). The calculated band gap (3.19 eV) shows excellent agreement with the experimental value (3.2 eV).5 Analysis of the projected density of states (Fig. 2) indicates that the valence band maximum (VBM) is dominated by O 2p states while conduction band minimum (CBM) is composed of Ti 3d states. We will now discuss the effect of doping and codoping on the electronic structure of SrTiO3.
image file: c4ra07289h-f1.tif
Fig. 1 Electronic band structure along the high symmetry k-path of the Brillouin zone for (a) undoped SrTiO3 (b) N-doped SrTiO3 (c) Sb-doped SrTiO3 (d) (Sb, N)-codoped SrTiO3. The horizontal dashed lines represent the Fermi level.

image file: c4ra07289h-f2.tif
Fig. 2 DOS and PDOS plots for undoped SrTiO3. The vertical dashed lines indicate the Fermi level.

3.2. N-doped SrTiO3

Since, doping of nitrogen in the oxide based material mainly involves replacement of oxygen by nitrogen, we consider only substitutional doping. From the optimized cell structure, obtained in our calculation, it is found that N doping does not make any significant change in the SrTiO3 crystal structure. The Ti–N bond length (1.976 Å) in N-doped SrTiO3 is slightly higher than the Ti–O bond length (1.974 Å) in undoped crystal. This leads to increase in lattice parameter by only a small extent.

Introduction of N (2s22p3) in place of O (2s22p4) is found to bring significant changes in the electronic structure of SrTiO3. The N-doped SrTiO3 is deficient by one electron. Comparison of band structure plot of the N-doped SrTiO3 (Fig. 1b) with that of the undoped SrTiO3 (Fig. 1a) shows that, some occupied and unoccupied states are introduced above the VB. Analysis of PDOS plot (Fig. 3a) indicates that these states are formed due to mixing of N 2p state with O 2p state along with small contribution from Ti 3d state. Consequently, the band gap (energy difference between the occupied impurity states and the conduction band) is reduced to 2.31 eV, which is responsible for the visible light absorption of N-doped SrTiO3. However, these localized states are highly undesirable for the photocatalytic purpose as they may hinder the mobility of the charge carriers. Therefore one needs to introduce a codopant to passivate these discrete states for better photocatalytic property. Since the CBM of SrTiO3 (prominent Ti 3d character), is located just 0.8 eV above the water reduction level (H+/H2), we carefully choose Sb as the codopant, which is expected not to perturb the conduction band (CB) level largely. Before discussing the results of the codoped system, it will be instructive to first discuss the monodoped system with Sb as the dopant.


image file: c4ra07289h-f3.tif
Fig. 3 DOS and PDOS plots for (a) N-doped SrTiO3 (b) Sb-doped SrTiO3. The vertical dashed lines indicate the Fermi level.

3.3. Sb-doped SrTiO3

Previous experimental reports show that, Sb occupies exclusively Ti lattice site in the Sb-doped SrTiO3 as the ionic size of Sb5+ (Ref = 0.60 Å) is closer to the ionic size of Ti4+ (Ref = 0.605 Å) than that of Sr (Ref = 1.44 Å).16,29,35,36,42 The optimized structure shows that the introduction of Sb into the Ti lattice site does not lead to much change in the lattice structure. As the stable oxidation state for Sb is the pentavalent state (Sb5+), it leaves one extra electron to the system. This is manifested in the band structure plot (Fig. 1c), with the Fermi level located in the conduction band. Analysis of PDOS plot (Fig. 3b) clearly indicates that the defects states are contributed by Ti 3d state. This is due to localization of the extra unpaired electron in the Ti 3d orbital. One more thing we should point out is that the respective band edges are still dominated by the O 2p and Ti 3d states, as observed in the case of undoped SrTiO3. The contribution of the Sb 5s and 5p states to the band edges is found to be less significant. This may be due to more ionic character of the Sb–O bond as indicated in the study of Wang et al.42 This is again supported by the analysis of Bader charge density.43 In the case of undoped SrTiO3, the calculated Bader charges on Ti and O centers are 2.04|e| and −1.23|e|, respectively, while, for Sb-doped SrTiO3, the values on the Sb and O centers are 2.80|e| and −1.29|e|, respectively.

3.4. (Sb, N)-codoped SrTiO3

Before going to the discussion on the electronic structure of (Sb, N)-codoped SrTiO3, we calculate the defect pair binding energy (Eb) using the relation44
 
Eb = ESb-SrTiO3 + EN-SrTiO3E(Sb,N-SrTiO3)ESrTiO3 (2)
where, EN-SrTiO3, ESb-SrTiO3, E(Sb,N)-SrTiO3 and ESrTiO3 represent the energy of the N-doped, Sb-doped, (Sb, N)-codoped, and undoped SrTiO3 supercell, respectively. The positive value of the defect pair binding energy (1.58 eV per supercell) indicates that the codoped system is sufficiently stable.

Let us discuss the synergistic effect of both N and Sb on the electronic structure of SrTiO3. The band structure plot (Fig. 1d) for the codoped SrTiO3 shows that the Fermi level resides above the VBM, similar to that of an intrinsic semiconductor. It is interesting to observe that the band gap reduces significantly to 2.66 eV, which ensures the absorption of visible light. This is the consequence of elevation of VB edge associated with the formation of (N 2p, O 2p) hybridized state (Fig. 4). The discrete acceptor states as appearing in the case of N-doped SrTiO3 are completely passivated, resulting into a continuum band structure. This is due to the presence of Sb which compensates for the electron deficiency introduced through N-doping. It has been quantified by Bader charge analysis, which indicates a higher electronic charge on the N centre in the (Sb, N)-codoped SrTiO3 (−1.36|e|) as compared to that in N-doped SrTiO3 (−1.15|e|). On the other hand, the Ti 3d defect states arising in the case of Sb-doped SrTiO3 is no longer present in the (Sb, N)-codoped SrTiO3. Presence of both Sb and N leads to formation of charge compensated system, which will minimize formation of undesired defects and thus reduce charge carrier loss.


image file: c4ra07289h-f4.tif
Fig. 4 DOS and PDOS plots for (Sb, N)-codoped SrTiO3. The vertical dashed lines indicate the Fermi level.

3.5. Defect formation energy

In order to find favorable growth condition for doping of the individual element as well as both, we calculate defect formation energy in each case.45,46 The defect formation energy relates chemical potential of the elements as
 
ΔHf = Edoped + nOμO + nTiμTiESrTiO3nNμNnSbμSb (3)
where, Edoped represents total energy for the doped/codoped SrTiO3 supercell. μ stands for chemical potential of the element. n is the number of elements inserted or removed for the construction of doped/codoped structure. At the equilibrium condition between the reservoir of Sr, Ti and O and SrTiO3 eqn (4) must be satisfied
 
μSr + μTi + 3μO = μSrTiO3(bulk) (4)

The chemical potential of the element cannot exceed that of the respect bulk (Sr/Ti) or gaseous (O) state, i.e. μSrμSr(bulk), μTiμTi(bulk) and μOμO(gas).

The expression for the heat of formation SrTiO3 is given by

 
Δ = μSrTiO3(bulk)μSr(bulk)μTi(bulk) − 3μO(gas) (5)

Since the heat of formation for SrTiO3 is a negative quantity we can write

 
μSr(bulk) + ΔμSrμSr(bulk) (6a)
 
μTi(bulk) + ΔμTiμTi(bulk) (6b)
 
3μO(gas) + Δ ≤ 3μO ≤ 3μO(gas) (6c)

The chemical potential, μSr(bulk), μTi(bulk), and μSb have been calculated from the energy of an atom in the respective bulk structure. On the other hand, μO(gas) and μN have been calculated from the energy of an oxygen atom (1/2EO2) and nitrogen atom (1/2EN2) in the corresponding gaseous molecule confined at the centre of a 20 × 20 × 20 Å3 cubic box.

The calculation of the formation energy (Fig. 5a and b) indicates that in both the cases substitutional doping is energetically more favored in comparison to the interstitial doping. Fig. 5a shows the variation of formation energy for N-doped SrTiO3 as a function of oxygen chemical potential (μOμO(gas)). Under O-rich condition the formation energy has a large positive value, which shows a decreasing trend as we proceed towards O-poor condition. As shown in Fig. 5b, that the formation energy for Sb-doping is highly positive under Ti-rich condition. This indicates that the doping with individual element is highly unfavorable under host rich condition, may be due to requirement of large energy for vacancy formation in the substitution process. However, this is found to be feasible in presence of both the dopant elements as shown in Fig. 5c. The calculation indicates that the O-rich condition is more suitable than the Ti-rich condition for the growth of the (Sb, N)-codoped SrTiO3.


image file: c4ra07289h-f5.tif
Fig. 5 The variation of defect formation energy with the chemical potential of O (μOμO-gas) and Ti (μTiμTi-bulk) for (a) N-doped SrTiO3 (b) Sb-doped SrTiO3 (c) (Sb, N)-codoped SrTiO3. The color lines (5c) corresponds to different formation energies for the (Sb, N)-codoped SrTiO3. The codoped system cannot be formed in the lower half of the diagonal axis (large white region) in the case of 5c because the chemical potential of Sr (μSr) is exceeded (eqn (6a)).

3.6. Optical property

To investigate the absorption property we calculate the frequency dependent dielectric function, ε(ω) = ε1(ω) + 2(ω). The real part (ε1) and imaginary part (ε2) of the dielectric tensor are obtained from Kramers–Kronig transformation, and using summation over the empty states, respectively, as implemented in VASP. The absorption coefficient α(ω) has been evaluated using the formula47
 
image file: c4ra07289h-t1.tif(7)

Fig. 6 shows that the absorption curve for the undoped SrTiO3 is limited to the UV region, while it is shifted towards the visible region in presence of both Sb and N dopants. This observation supports the band gap reduction of SrTiO3 due to codoping with Sb and N as discussed in the earlier section.


image file: c4ra07289h-f6.tif
Fig. 6 The calculated optical absorption curves for the undoped and codoped SrTiO3.

3.7. Photocatalytic activity

As we know, reduction in the band gap is necessary to improve the visible light activity, but not sufficient to achieve high photocatalytic activity for water splitting. The band edges must be in proper position with respect to the water redox levels for spontaneous release of hydrogen as well as oxygen from water splitting. The conduction band potential should be more positive than the H+/H2 potential and the valence band potential should be more negative than the H2O/O2 potential, i.e. CB should be located above the water redox potential (H+/H2) and VB should be located below the water oxidation level (H2O/O2). To check this criterion for our proposed system we align the positions of the band edges with respect to that of the undoped SrTiO3. For undoped SrTiO3 the CBM is located 0.8 eV above the water reduction (H+/H2) level and the VBM is 1.17 eV below the water oxidation level (H2O/O2).48 The VBM and CBM for the (Sb, N)-codoped SrTiO3 is now placed by considering the relative shifts of the corresponding electronic energy levels with respect to that of the undoped SrTiO3. As can be seen from Fig. 7, that the VBM of (Sb, N)-codoped SrTiO3 is located 0.20 eV below the water oxidation level and the CBM position is 1.23 eV above the water reduction level. Hence, overall water splitting is thermodynamically feasible in the (Sb, N)-codoped SrTiO3. Interestingly, the CBM of the (Sb, N)-SrTiO3 is located even 0.43 eV above the CBM of the undoped SrTiO3. This implies that the modified system should be more feasible for H2 evolution due to enhancement of reducing power with respect to the undoped SrTiO3.
image file: c4ra07289h-f7.tif
Fig. 7 The band alignment of the undoped, and codoped SrTiO3 with respect to the water redox levels.

3.8. Effect of supercell size

To investigate the effect of supercell size, we have also considered larger supercell (2 × 2 × 3). The formation energy has been calculated following the same procedure as discussed in the case of 2 × 2 × 2 supercell. Fig. 8 shows the formation energy variation as a function of chemical potential of Ti and O. The calculated formation energy for (Sb, N)-codoped SrTiO3 using 2 × 2 × 2 (Fig. 5c) and 2 × 2 × 3 supercell (Fig. 8) has been found to be almost similar. This justifies the choice of supercell size used in the calculation.
image file: c4ra07289h-f8.tif
Fig. 8 The variation of defect formation energy with the chemical potential of O (μOμO-gas) and Ti (μTiμTi-bulk) for (Sb, N)-codoped SrTiO3 using 2 × 2 × 3 supercell. The color lines corresponds to different formation energies for the (Sb, N)-codoped SrTiO3. The codoped system cannot be formed in the lower half of the diagonal axis (large white region) because the chemical potential of Sr (μSr) is exceeded (eqn (6a)).

4. Conclusion

The hybrid DFT calculations reported here reveal the significance of codoping of Sb and N into SrTiO3 for improving the photocatalytic activity under visible light. The positive defect pair binding energy indicates that the (Sb, N)-codoped SrTiO3 is sufficiently stable. The band gap is reduced significantly without introducing any isolated midgap states. As the codoped system is charge compensated, it is expected to reduce undesirable vacancy formation, and consequently defect assisted carrier loss can be avoided. Presence of Sb also facilitates the introduction of N into the SrTiO3 crystal structure by reducing formation energy. The band alignment for the codoped SrTiO3 is such that both the photo-oxidation and photo-reduction processes associated with water splitting are thermodynamically feasible. Moreover, the reducing power of the modified SrTiO3 is higher than that of the undoped material, making it more suitable for H2 generation. Hence we can propose that the codoping of Sb and N into the SrTiO3 crystal is one of the promising approaches to enhance the efficiency for photo-splitting of water using visible light.

Acknowledgements

We thank the BARC computer centre for providing the high performance parallel computing facility. The authors like to acknowledge Dr A. K. Samanta, Dr C. Majumder and Dr D. Karmakar for valuable discussions. We also thank Dr B. N. Jagatap for his encouragement and support. The work of Prof. Swapan K. Ghosh is supported through Sir J. C. Bose Fellowship from the Department of Science and Technology, India.

References

  1. A. Kudo and Y. Miseki, Chem. Soc. Rev., 2009, 38, 253–278 RSC.
  2. X. Chen, S. Shen, L. Guo and S. S. Mao, Chem. Rev., 2010, 110, 6503–6570 CrossRef CAS PubMed.
  3. K. Maeda and K. Domen, J. Phys. Chem. Lett., 2010, 1, 2655–2661 CrossRef CAS.
  4. T. Hisatomi, J. Kubota and K. Domen, Chem. Soc. Rev., 2014 10.1039/c3cs60378d.
  5. M. Cardona, Phys. Rev., 1965, 140, A651 CrossRef.
  6. H. Irie, Y. Maruyama and K. Hashimoto, J. Phys. Chem. C, 2007, 111, 1847–1852 CAS.
  7. J. W. Liu, G. Chen, Z. H. Li and Z. G. Zhang, J. Solid State Chem., 2006, 179, 3704–3708 CrossRef CAS PubMed.
  8. D. Wang, J. Ye, T. Kako and T. Kimura, J. Phys. Chem. B, 2006, 110, 15824–15830 CrossRef CAS PubMed.
  9. W. Wei, Y. Dai, H. Jin and B. Huang, J. Phys. D: Appl. Phys., 2009, 42, 055401 CrossRef.
  10. H. Yu, S. Ouyang, S. Yan, Z. Li, T. Yu and Z. Zou, J. Mater. Chem., 2011, 21, 11347–11351 RSC.
  11. P. Reunchan, N. Umezawa, S. Ouyang and J. Ye, Phys. Chem. Chem. Phys., 2012, 14, 1876–1880 RSC.
  12. R. Konta, T. Ishii, H. Kato and A. Kudo, J. Phys. Chem. B, 2004, 108, 8992–8995 CrossRef CAS.
  13. S. W. Bae, P. H. Borse and J. S. Lee, Appl. Phys. Lett., 2008, 92, 104107 CrossRef PubMed.
  14. H.-C. Chen, C.-W. Huang, J. C. S. Wu and S.-T. Lin, J. Phys. Chem. C, 2012, 116, 7897–7903 CAS.
  15. S. Kawasaki, K. Akagi, K. Nakatsuji, S. Yamamoto, I. Matsuda, Y. Harada, J. Yoshinobu, F. Komori, R. Takahashi, M. Lippmaa, C. Sakai, H. Niwa, M. Oshima, K. Iwashina and A. Kudo, J. Phys. Chem. C, 2012, 116, 24445–24448 CAS.
  16. K. Furuhashi, Q. Jia, A. Kudo and H. Onishi, J. Phys. Chem. C, 2013, 117, 19101–19106 CAS.
  17. J. Wang, S. Yin, M. Komatsu, Q. Zhang, F. Saito and T. Sato, J. Mater. Chem., 2003, 13, 2348–2352 RSC.
  18. H. W. Kang and S. B. Park, Chem. Eng. Sci., 2013, 100, 384–391 CrossRef CAS PubMed.
  19. N. Li and K. L. Yao, AIP Adv., 2012, 2, 032135 CrossRef PubMed.
  20. C. Zhang, Y. Jia, Y. Jing, Y. Yao, J. Ma and J. Sun, Comput. Mater. Sci., 2013, 79, 69–74 CrossRef CAS PubMed.
  21. J. Wang, S. Yin, M. Komatsu, Q. Zhang, F. Saito and T. Sato, J. Photochem. Photobiol., A, 2004, 165, 149–156 CrossRef CAS PubMed.
  22. Y. Y. Mi, S. J. Wang, J. W. Chai, J. S. Pan, C. H. A. Huan, Y. P. Feng and C. K. Ong, Appl. Phys. Lett., 2006, 89, 231922 CrossRef PubMed.
  23. P. Liu, J. Nisar, B. Pathak and R. Ahuja, Int. J. Hydrogen Energy, 2012, 37, 11611–11617 CrossRef CAS PubMed.
  24. M. Miyauchi, M. Takashio and H. Tobimatsu, Langmuir, 2004, 20, 232–236 CrossRef CAS.
  25. J. Wang, S. Yin, M. Komatsu and T. Sato, J. Eur. Ceram. Soc., 2005, 25, 3207–3212 CrossRef CAS PubMed.
  26. W. Wei, Y. Dai, M. Guo, L. Yu, H. Jin, S. Han and B. Huang, Phys. Chem. Chem. Phys., 2010, 12, 7612–7619 RSC.
  27. H. Yu, S. Yan, Z. Li, T. Yu and Z. Zou, Int. J. Hydrogen Energy, 2012, 37, 12120–12127 CrossRef CAS PubMed.
  28. R. D. Shannon, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr., 1976, 32, 751–767 CrossRef.
  29. H. Kato and A. Kudo, J. Phys. Chem. B, 2002, 106, 5029–5034 CrossRef CAS.
  30. R. Niishiro, R. Konta, H. Kato, W.-J. Chun, K. Asakura and A. Kudo, J. Phys. Chem. C, 2007, 111, 17420–17426 CAS.
  31. T. Ikeda, T. Nomoto, K. Eda, Y. Mizutani, H. Kato, A. Kudo and H. Onishi, J. Phys. Chem. C, 2008, 112, 1167–1173 CAS.
  32. C. Di Valentin, G. Pacchioni, H. Onishi and A. Kudo, Chem. Phys. Lett., 2009, 469, 166–171 CrossRef CAS PubMed.
  33. M. Niu, W. Xu, X. Shao and D. Cheng, Appl. Phys. Lett., 2011, 99, 203111–203113 CrossRef PubMed.
  34. P. Reunchan, S. Ouyang, N. Umezawa, H. Xu, Y. Zhang and J. Ye, J. Mater. Chem. A, 2013, 1, 4221–4227 CAS.
  35. R. Asai, H. Nemoto, Q. Jia, K. Saito, A. Iwase and A. Kudo, Chem. Commun., 2014, 50, 2543–2546 RSC.
  36. R. Niishiro, S. Tanaka and A. Kudo, Appl. Catal., B, 2014, 150–151, 187–196 CrossRef CAS PubMed.
  37. J. Paier, M. Marsman, K. Hummer, G. Kresse, I. C. Gerber and J. G. Ángyán, J. Chem. Phys., 2006, 124, 154709 CrossRef CAS PubMed.
  38. G. Kresse and D. Joubert, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758 CrossRef CAS.
  39. P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953 CrossRef.
  40. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188 CrossRef.
  41. J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J. Singh and C. Fiolhais, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 46, 6671 CrossRef CAS.
  42. H. H. Wang, F. Chen, S. Y. Dai, T. Zhao, H. B. Lu, D. F. Cui, Y. L. Zhou, Z. H. Chen and G. Z. Yang, Appl. Phys. Lett., 2001, 78, 1676–1678 CrossRef CAS PubMed.
  43. G. Henkelman, A. Arnaldsson and H. Jónsson, Comput. Mater. Sci., 2006, 36, 254–360 CrossRef PubMed.
  44. W.-J. Yin, S.-H. Wei, M. M. Al-Jassim and Y. Yan, Phys. Rev. Lett., 2011, 106, 066801 CrossRef.
  45. R. Long and N. J. English, Chem. Mater., 2010, 22, 1616–1623 CrossRef CAS.
  46. W.-J. Shi and S.-J. Xiong, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 205210 CrossRef.
  47. M. Fox, Optical Properties of Solids, Oxford University Press, USA, 2002 Search PubMed.
  48. Y. Xu and M. A. A. Schoonen, Am. Mineral., 2000, 85, 543 CAS.

This journal is © The Royal Society of Chemistry 2014