arxiv: v1 [cond-mat.mtrl-sci] 6 Sep 2011

Similar documents
From 180º stripe domains to more exotic patterns of polarization in ferroelectric nanostructures. A first principles view

Publication I. c 2010 American Physical Society. Reprinted with permission.

arxiv: v1 [cond-mat.mtrl-sci] 23 Oct 2013

Polarization Enhancement in Perovskite Superlattices by Oxygen. Octahedral Tilts

Effect of substrate-induced strains on the spontaneous polarization of epitaxial BiFeO 3 thin films

Supplementary Information for Dimensionality-Driven. Insulator-Metal Transition in A-site Excess. Nonstoichiometric Perovskites

Electronic Properties of Strained Si/Ge Core-Shell Nanowires. Xihong Peng, 1* Paul Logan 2 ABSTRACT

Phase Transitions in Strontium Titanate

Towards Direct-Gap Silicon Phases by the Inverse Band Structure Design Approach. P. R. China. Jiangsu , People s Republic of China

This is an electronic reprint of the original article. This reprint may differ from the original in pagination and typographic detail.

and strong interlayer quantum confinement

Supplementary Information

Supporting Information

First Principles Studies on the Electronic Structure and Band Structure of Paraelectric SrTiO 3 by Different Approximations

Exchange-induced negative-u charge order in N-doped WO 3 : A spin-peierls-like system

Origin of the Superior Conductivity of Perovskite Ba(Sr)SnO 3

Puckering and spin orbit interaction in nano-slabs

PBS: FROM SOLIDS TO CLUSTERS

Interfacial Coherency and Ferroelectricity of BaTiO 3 /SrTiO 3 Superlattice Films

Tinselenidene: a Two-dimensional Auxetic Material with Ultralow Lattice Thermal Conductivity and Ultrahigh Hole Mobility

What so special about LaAlO3/SrTiO3 interface? Magnetism, Superconductivity and their coexistence at the interface

Strain-induced single-domain growth of epitaxial SrRuO 3 layers on SrTiO 3 : a high-temperature x-ray diffraction study

University of Chinese Academy of Sciences, Beijing , People s Republic of China,

Topological band-order transition and quantum spin Hall edge engineering in functionalized X-Bi(111) (X = Ga, In, and Tl) bilayer

Stabilization of highly polar BiFeO 3 like structure: a new interface design route for enhanced ferroelectricity in artificial perovskite superlattice

Roger Johnson Structure and Dynamics: Displacive phase transition Lecture 9

arxiv: v1 [cond-mat.mtrl-sci] 16 Jul 2007

SUPPLEMENTARY MATERIAL

Ryan Hatcher and Chris Bowen. Samsung Advanced Logic Lab, Samsung Blvd Austin, Tx 78754

Jie Ma and Lin-Wang Wang Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA. Abstract

Nanoxide electronics

Supplementary Figure 1 Two-dimensional map of the spin-orbit coupling correction to the scalar-relativistic DFT/LDA band gap. The calculations were

Defects in TiO 2 Crystals

Supporting information Chemical Design and Example of Transparent Bipolar Semiconductors

Supporting Information: Local Electronic Structure of a Single-Layer. Porphyrin-Containing Covalent Organic Framework

Stability of Cubic Zirconia and of Stoichiometric Zirconia Nanoparticles

Structure of CoO(001) surface from DFT+U calculations

Room-temperature tunable microwave properties of strained SrTiO 3 films

PCCP PAPER. 1 Introduction TH Vladislav Borisov,* ab Sergey Ostanin b and Ingrid Mertig ab. View Article Online View Journal View Issue

Effects of biaxial strain on the electronic structures and band. topologies of group-v elemental monolayers

Phonon calculations with SCAN

Nanoxide electronics

Competing Ferroic Orders The magnetoelectric effect

Our first-principles calculations were performed using the Vienna Ab Initio Simulation

Supporting Information

arxiv:cond-mat/ v1 9 Aug 2006

Lecture contents. Stress and strain Deformation potential. NNSE 618 Lecture #23

Electronic structure of barium titanate : an abinitio DFT study. Hong-Jian Feng, Fa-Min Liu

A model for the direct-to-indirect band-gap transition in monolayer MoSe 2 under strain

First-principles study of BiScO 3 1Àx - PbTiO 3 x piezoelectric alloys

Correlation-Driven Charge Order at a Mott Insulator - Band Insulator Digital Interface

Mustafa Uludogan 1, Tahir Cagin, William A. Goddard, III Materials and Process Simulation Center, Caltech, Pasadena, CA 91125, U.S.A.

The Nature of the Interlayer Interaction in Bulk. and Few-Layer Phosphorus

PHASE-FIELD SIMULATION OF DOMAIN STRUCTURE EVOLUTION IN FERROELECTRIC THIN FILMS

Some surprising results of the Kohn-Sham Density Functional

Yuan Ping 1,2,3*, Robert J. Nielsen 1,2, William A. Goddard III 1,2*

Supporting Information Tuning Local Electronic Structure of Single Layer MoS2 through Defect Engineering

arxiv: v1 [cond-mat.mtrl-sci] 8 Aug 2016

Surface stress and relaxation in metals

InAs quantum dots: Predicted electronic structure of free-standing versus GaAs-embedded structures

Lecture 4: Band theory

Supplementary Figure 1: Projected density of states (DOS) of the d states for the four titanium ions in the SmSr superlattice (Ti 1 -Ti 4 as defined

Stability of a Ferroelectric Phase with Electrical Domains in Multilayers

Support Information. For. Theoretical study of water adsorption and dissociation on Ta 3 N 5 (100) surfaces

Monte Carlo Simulation of Ferroelectric Domain Structure: Electrostatic and Elastic Strain Energy Contributions

University of Munich, Theresienstr. 41, Munich, Germany and b Department of Physics, University of California,

Calculation and Analysis of the Dielectric Functions for BaTiO 3, PbTiO 3, and PbZrO 3

Coherent Lattice Vibrations in Mono- and Few-Layer. WSe 2. Supporting Information for. 749, Republic of Korea

Energy-Level Alignment at the Interface of Graphene Fluoride and Boron Nitride Monolayers: An Investigation by Many-Body Perturbation Theory

Materials 218/UCSB: Phase transitions and polar materials

Supporting Information. Designing a Lower Band Gap Bulk Ferroelectric Material with a Sizable. Polarization at Room Temperature

Defects in Semiconductors

arxiv: v1 [cond-mat.mtrl-sci] 22 Apr 2010

Ferroelectricity and Antiferroelectricity in Elemental Group-V Monolayer Materials

Influence of tetragonal distortion on the topological electronic structure. of the half-heusler compound LaPtBi from first principles

Institute of High Performance Computing, Agency for Science, Technology and Research, 1 Fusionopolis Way, #16-16 Connexis, Singapore

SnO 2 Physical and Chemical Properties due to the Impurity Doping

Explaining the apparent arbitrariness of the LDA-1/2 self-energy. correction method applied to purely covalent systems

Towards ferroelectrically-controlled magnetism: Magnetoelectric effect in Fe/BaTiO 3 multilayers

CITY UNIVERSITY OF HONG KONG. Theoretical Study of Electronic and Electrical Properties of Silicon Nanowires

A theoretical study of stability, electronic, and optical properties of GeC and SnC

Specific Heat of Cubic Phase of Protonic Conductor SrZrO 3

Rashba spin-orbit coupling in the oxide 2D structures: The KTaO 3 (001) Surface

STRUCTURAL AND MECHANICAL PROPERTIES OF AMORPHOUS SILICON: AB-INITIO AND CLASSICAL MOLECULAR DYNAMICS STUDY

Atomic Models for Anionic Ligand Passivation of Cation- Rich Surfaces of IV-VI, II-VI, and III-V Colloidal Quantum Dots

Atomic structure and stability of AlN 0001 and 0001 surfaces

2 ( º ) Intensity (a.u.) Supplementary Figure 1. Crystal structure for composition Bi0.75Pb0.25Fe0.7Mn0.05Ti0.25O3. Highresolution

Supplementary Information

Effect of substituting of S for O: The sulfide perovskite BaZrS 3 investigated with density functional theory

CHAPTER 6. ELECTRONIC AND MAGNETIC STRUCTURE OF ZINC-BLENDE TYPE CaX (X = P, As and Sb) COMPOUNDS

arxiv: v1 [cond-mat.str-el] 16 Jan 2015

SUPPLEMENTARY INFORMATION

Structural Calculations phase stability, surfaces, interfaces etc

SUPPLEMENTARY INFORMATION

Supporting Information for Interfacial Effects on. the Band Edges of Functionalized Si Surfaces in. Liquid Water

Landau-Ginzburg model for antiferroelectric phase transitions based on microscopic symmetry

A COMPUTATIONAL INVESTIGATION OF MIGRATION ENTHALPIES AND ELECTRONIC STRUCTURE IN SrFeO 3-δ

Electronic structure of nanocrystal quantum-dot quantum wells

Structure, energetics, and vibrational properties of Si-H bond dissociation in silicon

Interaction between a single-molecule

Transcription:

Band Gap and Edge Engineering via Ferroic Distortion and Anisotropic Strain: The Case of SrTiO 3 Robert F. Berger, 1 Craig J. Fennie, 2 and Jeffrey B. Neaton 1, 1 Molecular Foundry, Lawrence Berkeley National Laboratory, Berkeley, CA 2 School of Applied and Engineering Physics, Cornell University, Ithaca, NY (Dated: October 22, 2018) arxiv:1109.1271v1 [cond-mat.mtrl-sci] 6 Sep 2011 Theeffectsofferroic distortionandbiaxialstrainonthebandgapandbandedgesofsrtio 3 (STO) are calculated using density functional theory and many-body perturbation theory. Anisotropic strains are shown to reduce the gap by breaking degeneracies at the band edges. Ferroic distortions are shown to widen the gap by allowing new band edge orbital mixings. Compressive biaxial strains raise band edge energies, while tensile strains lower them. To reduce the STO gap, one must lower the symmetry from cubic while suppressing ferroic distortions. Our calculations indicate that for engineered orientation of the growth direction along [111], the STO gap can be controllably and considerably reduced at room temperature. PACS numbers: Within the perovskite family (ABO 3 ), a remarkable range of functional phenomena ferroelectricity, ferromagnetism, multiferroicity, and superconductivity, for example can be accessed through chemical substitution at the A and B sites. [1] In recent years, biaxial strains induced by epitaxy have provided another route to tune perovskite properties. [2] In a well-known example, both theory and experiment have shown that thin films of SrTiO 3 (STO), a prototypical perovskite insulator that takes up a centrosymmetric cubic structure (P m3m) under standard conditions, become polar and ferroelectric at room temperature for sufficiently large biaxial strains induced by coherent epitaxial growth. [3, 4] In STO, the large biaxial strains achievable via epitaxy provide access to a rich landscape of ferroic distortions, both polar ferroelectric and nonpolar antiferrodistortive, [3, 5, 6] that are known to alter both ferroelectric polarization and the low-frequency dielectric response. [5] In addition to its well-studied dielectric response, STO has also long been explored in solar water splitting applications due to its favorable flat-band potentials and stability against degradation in water. [7] However, its indirect optical gap (3.2 ev [8]) is too large for efficient light harvesting in the solar spectrum. Prior work has focused on chemical substitution as a means for reducing perovskite gaps, [9] but with unintended consequences for other properties. In this Letter, we use density functional theory (DFT) and beyond to examine a new route, via anisotropic strain and ferroic distortions, for altering the band gap and band edges of STO. We show that while anisotropic strain in the absence of ferroic distortions reduces the gap, ferroic distortions increase it. In addition, band edge energies, important for electro- and photocatalysis, can also be controllably increased or reduced relative to vacuum. The predicted trends can be fully understood with symmetry arguments, suggesting well-defined routes for tuning the electronic structures of perovskite oxides. For example, using intuition developedhere, weshowthatstrainedthin filmsofstogrown along an alternate orientation, [111], have considerably reduced gaps relative to bulk at room temperature. STO structural and electronic properties are obtained using DFT within the local density approximation (LDA), using the VASP package [10] and PAW potentials. [11] Electrons taken to be valence are 4s 2 4p 6 5s 2 of Sr, 3s 2 3p 6 4s 2 3d 2 of Ti, and 2s 2 2p 4 of O. A planewave cutoff of 500 ev is used throughout. At room temperature, STO takes up the cubic perovskite structure (Pm3m) with a five-atom primitive cell, which we treat with a Γ-centered 8 8 8 k-point mesh. Near 100 K, STO is known to undergo a transition to an antiferrodistortive phase in which Ti O octahedra rotate, doubling the unit cell and lowering symmetry to I4/mcm. [12] For such phases requiring supercells, we use proportionally fewer k-points. These parameters result in excellent convergence and good agreement with previous work. The DFT-LDA near-gap band structure of optimized cubic STO (a=3.864 Å) is shown in Fig. 1. The conduction band minimum (CBM) lies at Γ and consists almost entirely of Ti 3d states, while the valence band maximum (VBM) lies at R and consists almost entirely of O 2p states. As expected within DFT-LDA, the calculated indirect gap of 1.80 ev is far narrowerthan the experimental optical gap of 3.2 ev. [8] To obtain reliable band gaps and edges, we use many-body perturbation theory within the GW approximation [13], known to give quantitative band gaps and band edge energies for select semiconducting systems in the absence of strong excitonic effects. [14] Our computed G 0 W 0 gap of cubic STO (see [27]) is 3.82 ev, 0.6 ev larger than experiment. [28] This overestimate is consistent with past calculations for STO [15] and a closely-related system, TiO 2, where deviations are suggested to arise from excitonic and polaronic effects. [16] Overestimate notwithstanding, our G 0 W 0 band edge energies at equilibrium are in reasonable agreement with experiment. Experiments place the CBM and VBM 2.6

2 FIG. 1: (Left) DFT-LDA band structure of cubic STO. (Right) Schematic pictures of the highest occupied and lowest unoccupied states of STO as linear combinations of atomic orbitals. Ti atoms are shown in white, O atoms in gray. and 5.7 ev below the vacuum level, respectively,[17] compared with our G 0 W 0 calculated values of 2.48 and 6.30 ev (see [29]). Following previous calculations, [5, 6] we model biaxially strained thin films of STO by periodic calculations in which in-plane lattice parameters are fixed while the perpendicular axis and atomic positions are allowed to relax. Our structures have in-plane strains of up to ±4%, and the resulting lattice parameters (Supplemental Information) and structural phase diagram agree well with previous work. [3, 5, 6] This range of strain is somewhat optimistic experimentally, as strains approaching ±4% could limit the achievable film thickness, making bulk optical measurements difficult, and potentially introducing interfaces and quantum confinement effects. Nonetheless, perovskite substrates that would induce these strains are available, [18] and there are no fundamental limitations on stabilization of thin films on these substrates. For ultrathin films, our results here represent a quantitative baseline upon which future studies of quantum confinement and interface effects may be understood. We start with the most common growth orientation, [001], and compute electronic gaps and band edge energies for biaxial strains up to ±4%; these are shown in Fig. 2, initially for the paraelectric case in which the [001] axis is relaxed but further distortion of the internal coordinates is forbidden (P 4/mmm symmetry). Interestingly, biaxial strains tune both CBM and VBM by several tenths ofan ev, suggestingaroute to optimize eachband edge relative to the potentials of the water splitting halfreactions. From Fig. 2b, compressive biaxial strains raise band edge energies relative to vacuum, and tensile biaxial strains lower these energies. For paraelectric STO, the DFT-LDA gap narrows under both compressive and tensile strain. That the gap is non-monotonic about the cubic structure is expected on symmetry grounds (Fig. 2c). Cubic STO has O h point group symmetry, and both its CBM and VBM are threefold-degenerate. Biaxial reshaping of the unit cell lowers the point group symmetry to D 4h, splitting these degenerate triples into a spread of singles and pairs. As a result, the gap is narrower for tetragonal paraelectric STO than for cubic STO. Notably, our G 0 W 0 corrections to Kohn-Sham DFT- LDA eigenvalues (Fig. 2a,b, green) remain quite constant with changing biaxial strain. (A similar result was reported for silicon under isotropic pressure.[19]) The same is true of G 0 W 0 corrections to STO structures with ferroic distortions (Supplemental Information). This supports the premise that DFT-LDA, while insufficient for producing numerically accurate optical gaps, correctly tracks their trends. [19, 20] For the remainder of this paper, we restrict our focus to DFT-LDA results, at times assuming a constant correction of 1.4 ev the difference between DFT-LDA and room-temperature optical gaps of equilibrium STO to provide concrete estimates for gaps under biaxial strain. We turn now to the issue of how ferroic distortions affectthe STObandgap. Twoclassesofcommonlystudied ferroic distortions in perovskites are ferroelectric polar modes (FE, relative translation of the cation and anion sublattices) and antiferrodistortive modes (AFD, collective rotation of the Ti O octahedra). While neither distortion is present in unstrained STO at room temperature, both are accessible via strain, temperature, and applied fields. [3, 4, 12] To illustrate how these distortions might affect the band gap, we freeze in displacement patterns associated with the eigenvectors of lowlying IR-active phonon modes along the [001], [110], and [111] axes, and examine the DFT-LDA gap as a function of amplitude. We freeze in AFD modes by interpolating between cubic STO and optimized structures with AFD rotations about the [001], [110], and [111] axes. Computed gaps of these distorted STO structures are shown in Fig. 3. For modest changes in structural energy, both FE (Fig. 3a) and AFD (Fig. 3b) modes widen the gap by as much as 0.2 ev. This widening can be understood with symmetry arguments. As the symmetry is lowered under both types of distortion, conduction (primarily Ti 3d) and valence (primarily O 2p) states are allowed to mix and repel one another. Table I shows the irreducible representations of Ti 3d and O 2p states at Γ and R within the O h point group. By symmetry, the CBM of cubic STO cannot mix with O 2p states at Γ or R. In the presence of FE distortions, inversion symmetry is broken, the distinction between even and odd (g and u) parity is eliminated, and the CBM is allowed to mix with O 2p states. This mixing was previously illustrated by Wheeler et al. [21] for STO in terms of linear combinations of atomic orbitals. Similarly, in the presence of AFD rotations, doubling of the unit cell renders Γ and R equivalent, allowing the CBM to mix with O 2p states. Likewise, the VBM of STO can mix with Ti 3d states only in the presence of FE and/or AFD distortions. As Ti 3d and O 2p states mix, they widen the STO band

3 FIG. 2: a) Computed gaps and b) band edges of paraelectric, biaxially reshaped STO (P 4/mmm symmetry) under biaxial strain perpendicular to [001]. c) Schematic illustration of the trends in band edges as a function of strain. FIG. 3: Computed gaps (top) of STO structures in which a) FE polarization modes and b) AFD rotation modes are frozen in. Per-formula-unit structural energies relative to cubic STO are also shown (bottom). FE amplitudes are expressed as average relative translations of Ti and O atoms, while AFD amplitudes are expressed as average translations of O atoms. AFD rotations refer to [001] = a 0 a 0 c, [110] = a a c 0, and [111] = a a a in Glazer notation. [25] TABLE I: Irreducible representations of Ti 3d and O 2p states at Γ and R within O h point group symmetry. The CBM and VBM of cubic STO are shown in red. Atomic orbitals Irreducible representations Ti 3d at Γ t 2g, e g Ti 3d at R t 2g, e g O 2p at Γ t 1u, t 1u, t 2u O 2p at R t 1g, t 2g, e g, a 1g gap. While the magnitude of the widening depends on the geometry of a given distortion, this trend is general. Interestingly, AFD rotations can transform the STO gap from indirect to direct, as has been previously noted in both theory [22, 23] and experiment. [22, 24] Having established how anisotropic strain and ferroic distortions impact the STO electronic structure separately, we present in Fig. 4 the computed gaps of structures with ferroic distortions under biaxial strain(assuming a [001] growth direction). The energetic landscape of FIG. 4: a) Computed gaps of STO structures under biaxial strain perpendicular to [001], optimized with individual ferroic distortions. b) Gaps of fully optimized, strained STO structures, computed at zero temperature. ferroic distortions in ground-state STO has been studied extensively within DFT-LDA. [5, 6, 26] Consistent with this work, full relaxation causes FE and AFD to combine along [001] under compressive strain, and along [110] under tensile strain. All ferroic distortions, optimized separately (Fig. 4a), widen the gap relative to paraelectric STO. Full atomic relaxation (Fig. 4b) widens the gap even further. Using biaxial strain and associated ferroic distortions, we predict the optical gap can be tuned by nearly 20% under ±4% strains. As we have shown, the breaking of symmetry in STO has two manifestations with competing effects on the band gap biaxial reshaping of the unit cell that narrows the gap through broken degeneracies, and ferroic distortions that widen the gap through new orbital mixings. The data in Fig. 4b suggest that, in biaxially strained STO at low temperatures, ferroic distortions have the dominant effect on the gap. To engineer a narrower STO band gap, we must reshape the unit cell while suppressing ferroic distortions. One way to suppress ferroic distortions is to work at room temperature. Using a Landau- Ginzburg-Devonshire model with empirical parameters, Pertsev et al. computed a phase diagram of STO under [001] biaxial strain. [3] At 300 K, their model predicts that STO is paraelectric between 1.8% compressive and 1.5% tensile strain. While this provides a range of strain within which the STO gap can be narrowed, ferroic dis-

4 FIG. 5: Computed gaps of STO using free energy minimized structures at 300 K under biaxial strain perpendicular to [111]. All structures are paraelectric. tortions widen the gap outside the range. If, however, STO is grown epitaxially along an axis other than [001], the possibility exists for reshaping the unit cell without inducing ferroic distortions. Unlike growth along [001], which forces in-plane Ti O distances to be too short or too long for optimal bonding, a growth axis of [111] imposes no constraints on Ti O distances. We therefore expect that the [111] case does not strongly enhance ferroic distortions in STO. Using the same empirical parameters of Ref. [3], but applying them to films grown along [111] at 300 K, confirms that STO is expected to remain paraelectric over the entire range within ±4% strain. We therefore predict that [111] strain at room temperature (achievable via epitaxial growth on the (111) surfaces of perovskite substrates) could narrow the STO band gap significantly more than [001] strain (Fig. 5). Within DFT-LDA, 4% tensile strain perpendicular to [111] narrows the Kohn-Sham gap by more than 0.3 ev, to 1.46 ev. Using our constant correction, we therefore estimate that this approach could lower the optical gap of STO below 3 ev. To conclude, wehaveexamined the extent to which engineering of the STO crystal structure can tune its electronic gap and band edges. While biaxial strain widens the STO gap at zero temperature via ferroic distortions, it has the ability to narrow the gap significantly at room temperature, especially when applied perpendicular to the [111] direction. The observed trends can be understood on general symmetry grounds, suggesting that similar intuition can be productively applied across the broad family of perovskites. We are indebted to D. G. Schlom for inspiring this project and insightful discussions. This material is based upon work supported as part of the Energy Materials Center at Cornell (EMC 2 ), an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Award Number de-sc0001086. Work at the Molecular Foundry was supported by the Office of Science, Office of Basic Energy Sciences, of the U.S. Department of Energy under Contract Number DE-AC02-05CH11231. Electronic address: jbneaton@lbl.gov [1] T. Wolfram and S. Ellialtioglu, Electronic and Optical Properties of d-band Perovskites (Cambridge University Press, Cambridge, UK, 2006). [2] M. Dawber, K. M. Rabe, and J. F. Scott, Rev. Mod. Phys. 77, 1083 (2005). [3] N. A. Pertsev, A. K. Tagantsev, and N. Setter, Phys. Rev. B 61, R825 (2000). [4] J. H. Haeni et al., Nature 430, 758 (2004). [5] A. Antons, J. B. Neaton, K. M. Rabe, and D. Vanderbilt, Phys. Rev. B 71, 024102 (2005). [6] O. Diéguez, K. M. Rabe, and D. Vanderbilt, Phys. Rev. B 72, 144101 (2005); T. Hashimoto et al., Jpn. J. Appl. Phys.44, 7134(2005); C. H.Lin, C. M. Huang, andg.y. Guo, J. Appl. Phys. 100, 084104 (2006). [7] T. Bak, J. Nowotny, M. Rekas, and C. C. Sorrell, Int. J. Hydrogen Energ. 27, 991 (2002). [8] M. Cardona, Phys. Rev. 140, A651 (1965). [9] A. Fujimori et al., Phys. Rev. B 46, 9841 (1992); A. Fujimori, J. Phys. Chem. Solids 53, 1595 (1992); J. W. Bennett, I. Grinberg, and A. M. Rappe, J. Am. Chem. Soc. 130, 17409 (2008); J. W. Bennett, I. Grinberg, and A. M. Rappe, Phys. Rev. B 79, 235115 (2009). [10] G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169 (1996); J. P. Perdew and A. Zunger, Phys. Rev. B 23, 5048 (1981). [11] G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999). [12] P. A. Fleury, J. F. Scott, and J. M. Worlock, Phys. Rev. Lett. 21, 16 (1968); K. A. Müller, W. Berlinger, and F. Waldner, Phys. Rev. Lett. 21, 814 (1968). [13] M. Shishkin and G. Kresse, Phys. Rev. B 74, 035101 (2006). [14] M. S. Hybertsen and S. G. Louie, Phys. Rev. B 34, 5390 (1986). [15] G. Cappellini et al., J. Phys. Condens. Mat. 12, 3671 (2000); M. van Schilfgaarde, T. Kotani, and S. Faleev, Phys. Rev. Lett. 96, 226402 (2006); D. R. Hamann and D. Vanderbilt, Phys. Rev. B 79, 045109 (2009); C. Friedrich, S. Blügel, and A. Schindlmayr, Phys. Rev. B 81, 125102 (2010); M. S. Kim and C. H. Park, J. Korean Phys. Soc. 56, 490 (2010). [16] W. Kang and M. S. Hybertsen, Phys. Rev. B 82, 085203 (2010). [17] W. Maus-Friedrichs et al., Surf. Sci 515, 499 (2002). [18] D. G. Schlom et al., Annu. Rev. Mater. Res. 37, 589 (2007). [19] X. Zhu, S. Fahy, and S. G. Louie, Phys. Rev. B 39, 7840 (1989). [20] X. Zhao, C. M. Wei, L. Yang, and M. Y. Chou, Phys. Rev. Lett. 92, 236805 (2004); V. Barone et al., Nano Lett. 5, 1621 (2005). [21] R. A. Wheeler et al., J. Am. Chem. Soc. 108, 2222 (1986). [22] P. Galinetto et al., Ferroelectrics 337, 179 (2006). [23] E. Heifets, E. Kotomin, and V. A. Trepakov, J. Phys. Condens. Mat. 18, 4845 (2006). [24] S. I. Shablaev, A. M. Danishevskii, and V. K. Subashiev, Zh. Eksp. Teor. Fiz. 86, 2158 (1984). [25] A. M. Glazer, Acta Crystallogr. B28, 3384 (1972). [26] W. Zhong and D. Vanderbilt, Phys. Rev. Lett. 74, 2587 (1995); N. Sai and D. Vanderbilt, Phys. Rev. B 62, 13942

5 (2000). [27] G 0W 0 corrections to DFT-LDA eigenvalues are computed in VASP using a frequency-dependent RPA dielectric function sampling 96 frequency points, 288 bands (268 unfilled), a plane-wave cutoff for the response function of 300 ev, and a Γ-centered 6 6 6 k-point mesh. [28] In our STO calculations, 0.1 ev of the difference between G 0W 0 and experimental gaps is due to our use of the DFT-LDA optimized lattice parameter, which is more than 1% shorter than that of experiment. [29] Band edge energies are calculated in two steps. A slab of six perovskite layers with unrelaxed TiO 2-terminated (001) surfaces and 15 Å of vacuum is used to reference a bulk-like Ti 3s semicore state to the vacuum level. A periodic calculation is then used to compute band edge energies relative to the same semicore state.