Next Article in Journal
Pure, Size Tunable ZnO Nanocrystals Assembled into Large Area PMMA Layer as Efficient Catalyst
Next Article in Special Issue
In-Situ Synthesis of Nb2O5/g-C3N4 Heterostructures as Highly Efficient Photocatalysts for Molecular H2 Evolution under Solar Illumination
Previous Article in Journal
Must the Best Laboratory Prepared Catalyst Also Be the Best in an Operational Application?
Previous Article in Special Issue
Enhancement of Hydrogen Productions by Accelerating Electron-Transfers of Sulfur Defects in the CuS@CuGaS2 Heterojunction Photocatalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Band Gap Modulation of Tantalum(V) Perovskite Semiconductors by Anion Control

by
Young-Il Kim
1,2,* and
Patrick M. Woodward
2
1
Department of Chemistry, Yeungnam University, Gyeongsan 38541, Korea
2
Department of Chemistry and Biochemistry, The Ohio State University, Columbus, OH 43210, USA
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(2), 161; https://doi.org/10.3390/catal9020161
Submission received: 11 January 2019 / Revised: 1 February 2019 / Accepted: 3 February 2019 / Published: 7 February 2019
(This article belongs to the Special Issue Photocatalytic Hydrogen Evolution)

Abstract

:
Band gap magnitudes and valence band energies of Ta5+ containing simple perovskites (BaTaO2N, SrTaO2N, CaTaO2N, KTaO3, NaTaO3, and TaO2F) were studied by diffuse reflection absorbance measurements, density-functional theoretical calculations, and X-ray photoelectron spectroscopy. As a universal trend, the oxynitrides have wider valence bands and narrower band gaps than isostructural oxides, owing to the N 2p contribution to the electronic structure. Visible light-driven water splitting was achieved by using Pt-loaded CaTaO2N, together with a sacrificial agent CH3OH.

1. Introduction

Transition metal oxynitrides are of interest due to their potential as photocatalysts [1,2,3,4,5,6], pigments [7,8], battery electrodes [9], high-permittivity dielectrics [10], etc. Such a diverse functionality of oxynitrides is derived largely from the coexistence of O2−/N3− in the anion lattice. As is well established, the conduction and valence bands of simple perovskites AMX3 are based mostly on the frontier orbitals of M and X, respectively. In the case of oxide perovskites, O 2p orbitals participate in the valence band formation near the Fermi level. However, the inclusion of nitrogen, which brings a higher 2p orbital energy level than that of oxygen 2p, can effectively shift the top of the valence band upward resulting in the decreased band gap. It is interesting to note that if the energy difference between O 2p (−14.1 eV) and N 2p (−11.4 eV) orbitals [11] is reflected onto the valence band edge positions, many of the complex oxynitrides containing Ta5+, Nb5+, or Ti4+ would have band gaps falling in the visible light range (3.1~1.8 eV). Such a prospect in optical properties has motivated a number of studies on oxynitride perovskites and related phases, with views to semiconductor developments for visible light-harvesting photocatalysts or non-toxic inorganic pigments [1,2,3,4,5,6,7,8,12,13,14,15,16,17,18].
In 2001, Asahi et al. reported the visible light-driven photocatalytic activity of TiO2−xNx [1], which was followed by a number of studies on oxynitride-type photocatalysts. Promising photocatalysts were identified in various structure types such as simple perovskite (CaTaO2N, SrTaO2N, BaTaO2N, LaTiO2N, LaTaON2, CaNbO2N, SrNbO2N, BaNbO2N), complex perovskites (LaMgxTa1−xO1+3xN2−3x), spinel (ZnGa2OxNy), wurtzite (Ga1−xZnxN1−xOx), baddeleyite (TaON), and anosovite (Ta3N5) [3,4,5,6]. Among notable examples, Ga1−xZnxN1−xOx has an absorption edge at ~500 nm and showed a quantum yield of 5.2% for 410 nm light [19]. BaTaO2N-BaZrO3 solid solution could catalyze H2 evolution from water without sacrificial agents [20]. TaON showed overall water splitting activity with surface modification and appropriate co-catalysts [21]. LaMgxTa1−xO1+3xN2−3x and CaTaO2N could achieve overall water splitting with a Rh-Cr mixed oxide co-catalyst [22,23]. However, it was apparent that the photocatalytic behavior of a particular catalyst depended not merely on the composition but the morphology, defects, co-catalysts, and the type of photocatalytic reaction.
In this study, we compare the electronic structures of several Ta5+ perovskites having different anion matrices of pure oxide, oxynitride, and oxyfluoride types. The diffuse reflection absorbance spectra for ATaO2N (A = Ba, Sr, and Ca) are presented along with those of KTaO3, NaTaO3, and TaO2F, revealing a clear dependence of the semiconductor band gap on the electronegativity of anion components. The density-functional theory (DFT) based computations, combined with the valence level X-ray photoelectron spectroscopy (XPS), confirm that the N 2p component plays a critical role in extending the valence band edge in oxynitride compounds. We also present the photocatalytic activity of oxynitride samples tested by examining the water splitting under visible light irradiation.

2. Results and Discussion

Figure 1 displays the diffuse reflection absorbance spectra for simple Ta5+ perovskites where the oxynitride phases are found with markedly smaller band gap energies than the others. The optical band gaps were estimated by Shapiro’s method [24]. The linear region of the absorption edge was extrapolated to the wavelength axis, where the intersection (zero absorption) was taken as the band gap value. The estimated band gap energies are in the following order: BaTaO2N (1.8 eV) < SrTaO2N (2.1 eV) < CaTaO2N (2.4 eV) < KTaO3 (3.6 eV) < NaTaO3 (4.0 eV) < TaO2F (4.1 eV).
For the d0 perovskites AMX3, it has been well elucidated that the band gap magnitude depends on (i) electronegativity difference between M cation and X anion, (ii) deviation of MXM bond angles away from 180°, (iii) MX bond distance, and (iv) electronegativity of A cation [25,26]. The band gap variation among BaTaO2N (cubic), SrTaO2N (tetragonal), and CaTaO2N (orthorhombic) can be explained by the structural distortion factor (ii), in which the more distorted Ta−(O,N)−Ta linkage leads to the narrower band width and the wider band gap. However, the same reasoning cannot be used across distinct anion systems as the cubic KTaO3 has a greater band gap than that of CaTaO2N. In this regard, it can be judged that the control of anion components among N, O, and F, which have well-separated electronegativity values, makes a dominant effect on the resulting electronic structure. The absorbance spectra were also examined by using Tauc plots [27] from which the above oxynitride perovskites were found to be indirect-gap semiconductors.
A detailed aspect of the electronic structural evolution depending on the anion components was studied by band calculations at the DFT level and the XPS measurements. Structural parameters for DFT calculations were taken from the Rietveld refinements for BaTaO2N, SrTaO2N, and CaTaO2N [10], or from the literature data for KTaO3 [28], NaTaO3 [29], and TaO2F [30]. Since the computation codes cannot handle mixed occupation of any crystallographic site, ordered O/N (or O/F) distributions were assumed for mixed anion phases.
The density of states (DOS) in BaTaO2N, SrTaO2N, CaTaO2N, KTaO3, and TaO2F resulted from the calculations using CAmbridge Serial Total Energy Package (CASTEP) and are compared in Figure 2. As previously observed for similar compounds, the computation tends to underestimate the band gap magnitude. Still, it can be well recognized that the width and position of valence bands vary depending on the anion components. Both of the mixed anion systems have widened valence bands due to the 2p orbital mixings between O/N or O/F: extended toward a higher energy side for oxynitrides and toward a lower energy side for oxyfluoride. However, the conduction bands of those five compounds were found at fairly similar energy ranges (not shown) since they have the same octahedral cation, Ta. The net result is the effective band gap reduction in oxynitrides, as compared with oxides. On the other hand, for the oxyfluoride derivative, the band gap itself would not change very much to a first approximation.
The N 2p contribution to band structures of oxynitride compounds can be better viewed by extracting the partial DOS of the component atoms. Figure 3 shows the DOS plots for BaTaO2N as an example, which was obtained by employing linear muffin-tin orbital (LMTO) calculation. Both O p and N p orbitals were found as the major constituents of the valence band but notably the N character resides primarily at the upper region of the valence band, in agreement with the design concept of these oxynitride perovskites.
Along with the theoretical calculation, an experimental probe was also used to study the valence band structures of ATaO2N (A = Ba, Sr, Ca), KTaO3, NaTaO3, and TaO2F. The XPS spectra presented in Figure 4 were collected at near the Fermi level and, therefore, depict the DOS of valence bands. After the energy calibration using the C 1s peak energy and background subtraction, the tops of the valence bands were determined as indicated on the plots (Figure 4). The valence band edges of oxynitrides were found to be higher in energy by significant margins (≈1 eV) than the oxides’ or oxyfluoride’s, which is consistent with the electronic structure calculations. It is, therefore, corroborated both experimentally and theoretically that the hybridization of O 2p and N 2p orbitals are energetically feasible in the extended solid lattice, and that the partial N/O replacement can be a useful means to reduce the band gap size of oxide semiconductors. Based on the measured band gap magnitudes and the valence band widths, simplified band structures can be proposed for the Ta5+ perovskites studied here (Figure 5).
The reduced band gaps of oxynitride phases have immediate relevance to the photocatalytic reactivity. In this respect, we tested the water splitting by Pt-loaded oxynitride samples under visible light irradiation. Figure 6 presents the time-dependent H2 evolution from the Pt-CaTaO2N in H2O/CH3OH, along with the result from Pt-TiO2 (P25). Since CH3OH contains carbon with a formal oxidation number of −2, it can act as a reducing agent that removes O2 and expedite water decomposition as follows:
2H2O + hν → 2H2 + O2
2CH3OH + O2 → 2CO2 + 2H2
The sacrificial agent CH3OH should boost the generation of H2 according to the Le Chatelier principle, and also help suppress the H2−O2 recombination.
As displayed in Figure 6, Pt-CaTaO2N possesses the photocatalytic activity that can be triggered by visible light photons. Using the irradiation source of λ > 395 nm here, the photocatalytic efficiency of Pt-CaTaO2N is significantly higher than that of Pt-TiO2 (P25), a well-established photocatalyst system. Certainly, the superior performance of Pt-CaTaO2N is attributed to its narrower band gap. As can be found from Figure 1, CaTaO2N can utilize the photons with λ as long as ≈500 nm, whereas the absorption by TiO2 (P25) is limited to λ < 400 nm. The other oxynitride samples Pt-BaTaO2N and Pt-SrTaO2N and an oxide sample Pt-KTaO3 were also examined under the same experimental condition, but none of them produced discernible amounts of H2. The lack of photocatalytic ability in Pt-KTaO3 is simply ascribable to its wide band gap. However, in the cases of BaTaO2N and SrTaO2N, which possess even smaller band gap energies than CaTaO2N, the inferior photocatalytic property can be due to other factors. As one possibility, the valence band edges of BaTaO2N and SrTaO2N might be higher than the O2−/O2 oxidation level, or the H2−O2 recombination might occur so fast as to disallow the observation of water decomposition. Yet, BaTaO2N and SrTaO2N are regarded as promising candidates for visible light photocatalysts that could be well exploited in deliberately designed reaction systems. In the studies by Domen et al., it was demonstrated that the combination of Pt-ATaO2N (A = Ba, Sr, Ca) and Pt-WO3 achieves overall water splitting under visible light in the presence of IO3/I as a shuttle redox mediator [2,12].

3. Materials and Methods

3.1. Sample Syntheses and Crystal Structure

Polycrystalline oxynitride samples ATaO2N (A = Ba, Sr, Ca) were prepared by ammonolysis reaction using BaCO3 (J. T. Baker, 99.8%, Phillipsburg, NJ, USA), SrCO3 (Aldrich, 99.9+%, St. Louis, MO, USA), CaCO3 (Mallinckrodt, 99.95%, Phillipsburg, NJ, USA), and Ta2O5 (Cerac, 99.5%, Milwaukee, WI, USA), as described previously [10]. Quantitative mixture of reagents was heated in anhydrous ammonia (99.99%) at a flow rate of ≈50 cm3/min. Each ammonolytic heating cycle consisted of heating/cooling ramps of 10 °C/min and a dwell step of 20 h at 1000 °C. The heat treatment cycle was repeated 2–5 times to obtain phase pure products. For preparing TaO2F, Ta powder (Alfa Aesar, 99.9%, Karlsruhe, Germany) was dissolved in HF solution (47%) in a Teflon beaker. After evaporating the solvent at 125 °C, white precipitate was washed with distilled water, dried at 150 °C, and finally heated at 500 °C in air for 1 h. Reference compounds KTaO3 (Cerac, 99.9%) and NaTaO3 (Cerac, 99.9%) were used as purchased.
Crystal structure analyses of ATaO2N samples used the synchrotron X-ray powder diffraction patterns collected at the beamline X7A of National Synchrotron Light Source, Brookhaven National Laboratory (Upton, NY, USA). Lattice parameters and atomic coordinates for ATaO2N phases were refined using the Rietveld method as incorporated in the GSAS-GUI software suite [31,32].

3.2. Electronic Structure and Photocatalytic Property

Diffuse reflectance data were recorded and converted to absorbance using a spectrophotometer (Perkin Elmer, Lambda 20, Waltham, MA, USA) equipped with a 50-mm Labsphere integrating sphere over the spectral range 200–900 nm. The band gap energies were determined from Shapiro’s method [24] of extrapolating the onset of absorption to the wavelength axis.
DFT-based computations were performed using the CASTEP program as embodied in Accelrys Materials Studio [33]. Norm-conserving nonlocal pseudo-potentials were generated using the Kerker scheme with a kinetic energy cutoff of 400 eV. A convergence criterion of 0.02 meV was applied for the energy change per atom. Electron exchange and correlation were described using the Perdew-Wang generalized gradient approximation (PW91-GGA) [34]. For BaTaO2N, the total and partial densities of states were also calculated using a computation code, Stuttgart LMTO version 47, developed by Anderson and co-workers [35,36]. The program employs a TB-LMTO-ASA (tight binding linear muffin-tin orbital atomic sphere approximation) algorithm. Integrations over k space were performed using the tetrahedron method with a total of 40 irreducible k points from a 6 × 6 × 6 grid of reducible k points.
Valence band structures of ATaO2N (A = Ba, Sr, Ca), KTaO3, NaTaO3, and TaO2F were experimentally studied by XPS at near Fermi energy level, using a V. G. Scientific spectrometer equipped with a Mg Kα source (1253.6 eV) and operated at 9 kV and 20 mA with a base pressure of ≈2 × 10−9 Torr. Shirley method [37] was used for the data smoothening and background removal from the raw XPS spectra.
Photocatalytic activity of CaTaO2N, in comparison with that of TiO2 (Degussa P25) [38], was examined for the water decomposition using visible light irradiation. To focus on the photocatalytic H2 evolution, Pt was employed as a co-catalyst [39]. For preparing the Pt-impregnated catalyst, sample powder was stirred in an aqueous solution of H2PtCl6·6H2O ([Pt4+] ≈ 0.4 mM) under ultraviolet (UV) irradiation for 24 h, rinsed, and dried at room temperature. Thus, the obtained Pt-loaded catalyst (≈50 mg) was suspended in a mixture of 35 mL H2O and 0.6 mL MeOH contained in a 43.5 mL quartz vessel, which was sealed with a latex septum and filled with ≈1 atm of Ar. The photocatalytic reaction was induced by external illumination with an Oriel Xe lamp (24 V, 7 A) through a liquid filter and a long-pass filter (λcutoff = 395 nm), and was monitored using a gas chromatograph (Shimadzu, GC-14A, Tokyo, Japan) with Ar (99.998%) carrier gas. By using the liquid filter with a circulating water cooler, the reaction vessel was kept from the heating effect of infrared light component.

4. Conclusions

It was shown, using six simple perovskites with octahedral Ta5+, that the semiconductor band gap can be widely modulated by the electronegativity of anion components. The band gap generally widens from oxynitrides to oxides to oxyfluorides, and in most cases, the d0 oxynitride phases have band gaps corresponding to visible light energy. Band structure calculations by the DFT method and XPS measurements indicate that the N 2p component contributes to extend the top of the valence band in oxynitrides, making a principal distinction from the oxides’ electronic structures. The photocatalytic H2 generation from H2O was observed by using Pt-CaTaO2N and a sacrificial electron donor CH3OH under visible light.

Author Contributions

P.M.W. designed the research and reviewed the draft. Y.-I.K. prepared and characterized the samples and wrote the draft.

Funding

This research was funded by the National Research Foundation of Korea (NRF-2015R1D1A1A01056591) through the Basic Science Research Program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef] [PubMed]
  2. Higashi, M.; Abe, R.; Takata, T.; Domen, K. Photocatalytic Overall Water Splitting under Visible Light Using ATaO2N (A = Ca, Sr, Ba) and WO3 in a IO3/I Shuttle Redox Mediated System. Chem. Mater. 2009, 21, 1543–1549. [Google Scholar] [CrossRef]
  3. Takata, T.; Pan, C.; Domen, K. Recent progress in oxynitride photocatalysts for visible-light-driven water splitting. Sci. Technol. Adv. Mater. 2015, 16, 033506. [Google Scholar] [CrossRef] [PubMed]
  4. Wu, Y.; Lazic, P.; Hautier, G.; Persson, K.; Ceder, G. First principles high throughput screening of oxynitrides for water-splitting photocatalysts. Energy Environ. Sci. 2013, 6, 157–168. [Google Scholar] [CrossRef]
  5. Zhang, P.; Zhang, J.; Gong, J. Tantalum-based semiconductors for solar water splitting. Chem. Soc. Rev. 2014, 43, 4395–4422. [Google Scholar] [CrossRef] [PubMed]
  6. Ahmed, M.; Xinxin, G. A review of metal oxynitrides for photocatalysis. Inorg. Chem. Front. 2016, 3, 578–590. [Google Scholar] [CrossRef]
  7. Jansen, M.; Letschert, H.P. Inorganic yellow-red pigments without toxic metals. Nature 2000, 404, 980–982. [Google Scholar] [CrossRef]
  8. Tessier, F.; Marchand, R. Ternary and higher order rare-earth nitride materials: Synthesis and characterization of ionic-covalent oxynitride powders. J. Solid State Chem. 2003, 171, 143–151. [Google Scholar] [CrossRef]
  9. Cabana, J.; Dupre, N.; Gillot, F.; Chadwick, A.V.; Grey, C.P.; Palacin, M.R. Synthesis, Short-Range Structure, and Electrochemical Properties of New Phases in the Li−Mn−N−O System. Inorg. Chem. 2009, 48, 5141–5153. [Google Scholar] [CrossRef]
  10. Kim, Y.-I.; Woodward, P.M.; Baba-Kishi, K.Z.; Tai, C.W. Characterization of the Structural, Optical, and Dielectric Properties of Oxynitride Perovskites AMO2N (A = Ba, Sr, Ca; M = Ta, Nb). Chem. Mater. 2004, 16, 1267–1276. [Google Scholar] [CrossRef]
  11. Harrison, W.A. Electronic Structure and the Properties of Solids; W. H. Freeman and Company: San Francisco, CA, USA, 1980; p. 50. [Google Scholar]
  12. Higashi, M.; Abe, R.; Teramura, K.; Takata, T.; Ohtani, B.; Domen, K. Two step water splitting into H2 and O2 under visible light by ATaO2N (A = Ca, Sr, Ba) and WO3 with IO3/I shuttle redox mediator. Chem. Phys. Lett. 2008, 452, 120–123. [Google Scholar] [CrossRef]
  13. Yashima, M.; Yamada, H.; Maeda, K.; Domen, K. Experimental visualization of covalent bonds and structrural disorder in a gallium zince oxynitride photocatalyst (Ga1−xZnx)(N1−xOx): Origin of visible light absorption. Chem. Commun. 2010, 46, 2379–2382. [Google Scholar] [CrossRef] [PubMed]
  14. Hisatomi, T.; Katayama, C.; Moriya, Y.; Minegishi, T.; Katayama, M.; Nishiyama, H.; Yamada, T.; Domen, K. Photocatalytic oxygen evolution using BaNbO2N modified with cobalt oxide under photoexcitation up to 740 nm. Energy Environ. Sci. 2013, 6, 3595–3599. [Google Scholar] [CrossRef]
  15. Xiao, M.; Wang, S.; Thaweesak, S.; Luo, B.; Wang, L. Tantalum (Oxy)Nitride: Narrow Bandgap Photocatalysts for Solar Hydrogen generation. Engineering 2017, 3, 365–378. [Google Scholar] [CrossRef]
  16. Mukherji, A.; Seger, B.; Lu, G.Q.; Wang, L. Nitrogen Doped Sr2Ta2O7 Coupled with Graphene Sheets as Photocatalysts for Increased Photocatalytic Hydrogen Production. ACS Nano 2011, 5, 3483–3492. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, S.; Yang, J.; Ding, C.; Li, R.; Jin, S.; Wang, D.; Han, H.; Zhang, F.; Li, C. Nitrogen-doped layered oxide Sr5Ta4O15−xNx for water reduction and oxidation under visible light irradiation. J. Mater. Chem. A 2013, 1, 5651–5659. [Google Scholar] [CrossRef]
  18. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef] [PubMed]
  19. Kubota, J.; Domen, K. Photocatalytic Water Splitting Using Oxynitride and Nitride Semiconductor Powders for Production of Solar Hydrogen. Electrochem. Soc. Interface 2013, 22, 57–62. [Google Scholar] [CrossRef]
  20. Matoba, T.; Maeda, K.; Domen, K. Activation of BaTaO2N photocatalyst for enhanced non-sacrificial hydrogen evolution from water under visible light by forming a solid solution with BaZrO3. Chem. Eur. J. 2011, 17, 14731–14735. [Google Scholar] [CrossRef] [PubMed]
  21. Maeda, K.; Lu, D.; Domen, K. Direct water splitting into hydrogen and oxygen under visible light by using modified TaON photocatalysts with d0 electronic configuration. Chem. Eur. J. 2013, 19, 4986–4991. [Google Scholar] [CrossRef] [PubMed]
  22. Pan, C.; Takata, T.; Nakabayashi, M.; Matsumoto, T.; Shibata, N.; Ikuhara, Y.; Domen, K. A complex perovskite-type oxynitride: The first photocatalyst for water splitting operable at up to 600 nm. Angew. Chem. Int. Ed. 2015, 54, 2955–2959. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, J.; Pan, C.; Takata, T.; Domen, K. Photocatalytic overall water splitting on the perovskite-type transition metal oxynitride CaTaO2N under visible light irradiation. Chem. Commun. 2015, 51, 7191–7194. [Google Scholar] [CrossRef]
  24. Shapiro, I.P. Determination of the forbidden zone width from diffuse reflection spectra. Opt. Specktroskopiya 1958, 4, 256–260. [Google Scholar]
  25. Cox, P.A. The Electronic Structure and Chemistry of Solids; Oxford University Press: Oxford, UK, 1987; pp. 45–72. [Google Scholar]
  26. Eng, H.W.; Barnes, P.W.; Auer, B.M.; Woodward, P.M. Investigations of the electronic structure of d0 transition metal oxides belonging to the perovskite family. J. Solid State Chem. 2003, 175, 94–109. [Google Scholar] [CrossRef]
  27. Tauc, J. Optical properties and electronic structure of amorphous Ge and Si. Mater. Res. Bull. 1968, 3, 37–46. [Google Scholar] [CrossRef]
  28. Zhurova, E.A.; Ivanov, Y.; Zavodnik, V.; Tsirelson, V. Electron density and atomic displacements in KTaO3. Acta Cryst. B 2000, 56, 594–600. [Google Scholar] [CrossRef]
  29. Kennedy, B.J.; Prodjosantoso, A.K.; Howard, C.J. Powder neutron diffraction study of the high temperature phase transitions in NaTaO3. J. Phys. Condens. Matter 1999, 11, 6319–6327. [Google Scholar] [CrossRef]
  30. Frevel, L.K.; Rinn, H.W. The crystal structure of NbO2F and TaO2F. Acta Cryst. 1956, 9, 626–627. [Google Scholar] [CrossRef]
  31. Larson, A.C.; von Dreele, R.B. General Structure Analysis System (GSAS)—Report LAUR 86-748; Los Alamos National Laboratory: Los Alamos, NM, USA, 1994.
  32. Toby, B. EXPGUI, a graphical user interface for GSAS. J. Appl. Cryst. 2001, 34, 210–213. [Google Scholar] [CrossRef]
  33. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys. Condens. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  34. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671–6687. [Google Scholar] [CrossRef]
  35. Anderson, O.K.; Jepsen, O. Explicit, First-Principles Tight-Binding Theory. Phys. Rev. Lett. 1984, 53, 2571–2574. [Google Scholar] [CrossRef]
  36. Anderson, O.K.; Pawlowska, Z.; Jepsen, O. Illustration of the linear-muffin-tin-orbital tight-binding representation: Compact orbitals and charge density in Si. Phys. Rev. B 1986, 34, 5253–5269. [Google Scholar] [CrossRef]
  37. Shirley, D.A. High-Resolution X-Ray Photoemission Spectrum of the Valence Bands of Gold. Phys. Rev. B 1972, 5, 4709–4714. [Google Scholar] [CrossRef]
  38. Ohtani, B.; Prieto-Mahaney, O.O.; Li, D.; Abe, R. What is Degussa (Evonik) P25? Crystalline composition analysis, reconstruction from isolated pure particles and photocatalytic activity test. J. Photochem. Photobiol. A Chem. 2010, 216, 179–182. [Google Scholar] [CrossRef]
  39. Yang, J.Y.; Wang, D.; Han, H.; Li, C. Roles of cocatalysts in photocatalysis and photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Diffuse reflection absorption spectra for (i) BaTaO2N, (ii) SrTaO2N, (iii) CaTaO2N, (iv) KTaO3, (v) NaTaO3, and (vi) TaO2F.
Figure 1. Diffuse reflection absorption spectra for (i) BaTaO2N, (ii) SrTaO2N, (iii) CaTaO2N, (iv) KTaO3, (v) NaTaO3, and (vi) TaO2F.
Catalysts 09 00161 g001
Figure 2. Valence band density of states (DOS) structures for BaTaO2N, SrTaO2N, CaTaO2N, KTaO3, and TaO2F as calculated using the CAmbridge Serial Total Energy Package (CASTEP) code.
Figure 2. Valence band density of states (DOS) structures for BaTaO2N, SrTaO2N, CaTaO2N, KTaO3, and TaO2F as calculated using the CAmbridge Serial Total Energy Package (CASTEP) code.
Catalysts 09 00161 g002
Figure 3. Total and partial DOS for BaTaO2N as calculated using linear muffin-tin orbital (LMTO) code.
Figure 3. Total and partial DOS for BaTaO2N as calculated using linear muffin-tin orbital (LMTO) code.
Catalysts 09 00161 g003
Figure 4. Valence level XPS spectra for (i) BaTaO2N, (ii) SrTaO2N, (iii) CaTaO2N, (iv) KTaO3, (v) NaTaO3, and (vi) TaO2F. Vertical bar on each data indicates the top edge of the valence band.
Figure 4. Valence level XPS spectra for (i) BaTaO2N, (ii) SrTaO2N, (iii) CaTaO2N, (iv) KTaO3, (v) NaTaO3, and (vi) TaO2F. Vertical bar on each data indicates the top edge of the valence band.
Catalysts 09 00161 g004
Figure 5. Conduction (unfilled) and valence (shaded) band positions for several simple perovskites with octahedral Ta5+, as deduced from diffuse reflection absorbance and XPS measurements.
Figure 5. Conduction (unfilled) and valence (shaded) band positions for several simple perovskites with octahedral Ta5+, as deduced from diffuse reflection absorbance and XPS measurements.
Catalysts 09 00161 g005
Figure 6. Photocatalytic H2 evolutions over Pt-loaded powders of CaTaO2N (filled circles) and P25 TiO2 (open circles). At the beginning and after 30 h had elapsed, the reactor vessel was purged with Ar.
Figure 6. Photocatalytic H2 evolutions over Pt-loaded powders of CaTaO2N (filled circles) and P25 TiO2 (open circles). At the beginning and after 30 h had elapsed, the reactor vessel was purged with Ar.
Catalysts 09 00161 g006

Share and Cite

MDPI and ACS Style

Kim, Y.-I.; Woodward, P.M. Band Gap Modulation of Tantalum(V) Perovskite Semiconductors by Anion Control. Catalysts 2019, 9, 161. https://doi.org/10.3390/catal9020161

AMA Style

Kim Y-I, Woodward PM. Band Gap Modulation of Tantalum(V) Perovskite Semiconductors by Anion Control. Catalysts. 2019; 9(2):161. https://doi.org/10.3390/catal9020161

Chicago/Turabian Style

Kim, Young-Il, and Patrick M. Woodward. 2019. "Band Gap Modulation of Tantalum(V) Perovskite Semiconductors by Anion Control" Catalysts 9, no. 2: 161. https://doi.org/10.3390/catal9020161

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop