Next Article in Journal
Structural and Electrochemical Behaviors of ZnO Structure: Effect of Different Zinc Precursor Molarity
Previous Article in Journal
Physicochemical Properties of Ti3+ Self-Doped TiO2 Loaded on Recycled Fly-Ash Based Zeolites for Degradation of Methyl Orange
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Systematic Trends in Hybrid-DFT Computations of BaTiO3/SrTiO3, PbTiO3/SrTiO3 and PbZrO3/SrZrO3 (001) Hetero Structures

1
Institute of Solid State Physics, University of Latvia, 8 Kengaraga Str., LV1063 Riga, Latvia
2
Transport and Telecommunication Institute, Lomonosov Str. 1, LV1019 Riga, Latvia
3
Laboratory of Theoretical and Computational Chemistry, Institute of Theoretical Chemistry, Jilin University, Changchun 130023, China
*
Author to whom correspondence should be addressed.
Condens. Matter 2022, 7(4), 70; https://doi.org/10.3390/condmat7040070
Submission received: 1 November 2022 / Revised: 21 November 2022 / Accepted: 25 November 2022 / Published: 29 November 2022

Abstract

:
We performed predictive hybrid-DFT computations for PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 (001) surfaces, as well as their BaTiO3/SrTiO3, PbTiO3/SrTiO3 and PbZrO3/SrZrO3 (001) heterostructures. According to our hybrid-DFT computations for BO2 and AO-terminated ABO3 solid (001) surfaces, in most cases, the upper layer ions relax inwards, whereas the second layer ions shift upwards. Our hybrid-DFT computed surface rumpling s for the BO2-terminated ABO3 perovskite (001) surfaces almost always is positive and is in a fair agreement with the available LEED and RHEED experiments. Computed B-O atom chemical bond population values in the ABO3 perovskite bulk are enhanced on its BO2-terminated (001) surfaces. Computed surface energies for BO2 and AO-terminated ABO3 perovskite (001) surfaces are comparable; thus, both (001) surface terminations may co-exist. Our computed ABO3 perovskite bulk Γ-Γ band gaps are in fair agreement with available experimental data. BO2 and AO-terminated (001) surface Γ-Γ band gaps are always reduced with regard to the respective bulk band gaps. For our computed BTO/STO and PTO/STO (001) interfaces, the average augmented upper-layer atom relaxation magnitudes increased by the number of augmented BTO or PTO (001) layers and always were stronger for TiO2-terminated than for BaO or PbO-terminated upper layers. Our B3PW concluded that BTO/STO, as well as SZO/PZO (001) interface Γ-Γ band gaps, very strongly depends on the upper augmented layer BO2 or AO-termination but considerably less so on the number of augmented (001) layers.

1. Introduction

Different aspects of (001) surface as well as (001) interface phenomena in ABO3 perovskites are topics of paramount importance in modern solid state physics [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16]. All of our hybrid-DFT computed SrZrO3, PbZrO3, PbTiO3, BaTiO3 and SrTiO3 solids belong to the group of so-called ABO3 perovskite compounds [17,18,19]. Strontium zirconate (SrZrO3) is an extremely interesting material due to its high-temperature electronic properties [20]. The SrZrO3 phase transitions have been very carefully examined by Kennedy et al. [21] by means of powder neutron diffraction, as well as the Rietveld method [21]. Maniwa et al. [22] and Shirane et al. [23] described PbZrO3 as antiferroelectric due to its dielectric behaviour. Since then, materials that revealed a structural phase transition between two non-polar phases with a large dielectric anomaly at the high-temperature region of the transition were called antiferroelectrics [22,23]. Antiferroelectric oxides, for example PbZrO3, become highly promising candidate materials for dielectric energy storage capacitors [24,25]. PbTiO3 is a crucial component in a wide range of industrially important applications, including sensors, ultrasonic transducers and actuators [26]. Polymer ceramic hybrid composites (PVA/PbTiO3) are highly promising functional materials in a wide variety of research disciplines, demonstrating useful electrical, optical, mechanical, thermal and antibacterial properties [27]. Ferroelectric materials, for example BaTiO3, have been mentioned in connection to bacterial degradation [28]. It is demonstrated that the remnant polarization can induce antibacterial activities [29] in ferroelectric materials like BaTiO3. Lead zirconate titanate Pb(Zr,Ti)O3 is the most widely used worldwide piezoceramic material for electromechanical device applications [30]. Strontium titanate exhibits photocatalytic activity [31,32]. The identification of highly promising thermoelectric materials, like SrTiO3, is accelerated by theoretical predictions from ab initio calculations [33,34,35,36]. In general, there is a lot of interest in the science and technology of transition metal oxides, like ABO3 perovskites, due to their large variety of specific electronic, optical and magnetic properties [37,38,39,40,41,42,43,44,45,46,47,48]. For that reason, during the last 25 years, a lot of theoretical and experimental studies have been performed for PbTiO3, BaTiO3, SrTiO3, SrZrO3 and PbZrO3 perovskite (001) surfaces [49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72]. At the same time, it is very difficult to understand why a smaller number of first-principles computations, as well as experimental studies, were devoted to BTO/STO and PTO/STO, as well as SZO/PZO (001) heterostructures [73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88].
SrTiO3 (STO) perovskite is paraelectric and maintains its cubic centrosymmetric structure at all observed temperatures [89]. BaTiO3 (BTO), as a function of temperature, undergoes three phase transitions. Namely, the BTO crystal structure changes at 183 K from rhombohedral (R3m) to orthorhombic (Amm2), then at 278 K to tetragonal (P4mm), and finally at 403 K to a high-temperature cubic structure ( P m 3 ¯ m ) [90]. PbTiO3 (PTO) undergoes only one phase transition at 766 K from a tetragonal ferroelectric phase, stable at room temperature, with the space group P4mm to the high-symmetry cubic structure with the space group P m 3 ¯ m [91]. It is well-known that the classical PbZrO3 (PZO) perovskite exhibits three different phases. Namely, at low temperatures, PZO has an orthorhombic antiferroelectric phase stable at the temperature region up to 230 °C. The PZO rhombohedral ferroelectric phase is stable at the temperature region from 230 °C to 233 °C. Finally, PZO perovskite possesses a cubic paraelectric phase at temperatures above 233 °C [69]. Lastly, for the SrZrO3 (SZO) perovskite, Ligny et al. [92] discovered the following three phase transitions from enthalphy measurements: orthorhombic (Pnma) → orthorhombic (Cmcm) → tetragonal (I4/mcm) → cubic ( P m 3 ¯ m ) at temperatures equal to 995 K and 1105 K, as well as 1140 K, respectively [92].
For our hybrid-DFT computations, we adopted the classical cubic unit cells of SrZrO3, PbZrO3, PbTiO3, BaTiO3 and SrTiO3 perovskites. The cubic unit cells of all five of our ab initio computed ABO3 perovskites contain five atoms [93,94,95,96]. They all have exactly the same space group P m 3 ¯ m , as well as the same space group number, 221. In the cubic ABO3 perovskite matrix, the A-type atom (Sr, Pb or Ba) was located at the cube corner position. Thereby, the ABO3 perovskite A atom, in our computations, is located at the coordinate system origin and has the following coordinates (0, 0, 0). The B-type ABO3 perovskite atom (Ti or Zr) is located at the body center location of the cube. The coordinates of the B atom are equal to (½, ½, ½). Finally, the ABO3 perovskite unit cell’s three oxygen atoms are located at the cube face center positions with the following coordinates (½, ½, 0), (½, 0, ½), (0, ½, ½).
The aim of our research reported herein was to perform necessary additional hybrid-DFT computations for SrZrO3, PbZrO3, PbTiO3, BaTiO3 and SrTiO3 perovskite (001) surfaces, as well as their (001) heterostructures, in order to complete our many years of lasting work in this direction, and to create a theory that explains systematic trends in ABO3 perovskite (001) surface and interface computations. In the first part of our contribution, we present comparative hybrid-DFT computation results dealing with the pristine SrZrO3, PbZrO3, PbTiO3, BaTiO3 and SrTiO3 (001) surfaces. We computed the SrZrO3, PbZrO3, PbTiO3, BaTiO3 and SrTiO3 (001) surface atom relaxations for both possible BO2 and AO (001) surface terminations and compared the calculated surface rumplings s and interlayer distances Δdij with the available experimental data. As a next step, we computed the SrZrO3, PbZrO3, PbTiO3, BaTiO3 and SrTiO3 (001) surface energies, as well as B-O chemical band populations near the BO2-terminated (001) surfaces. Finally, we computed the band gaps for bulk PbTiO3, BaTiO3 and SrTiO3 and their (001) surfaces by means of different exchange-correlation functionals. As a logical next step, we extended our (001) surface calculations to hybrid-DFT computations of considerably less-studied SrTiO3/BaTiO3, SrZrO3/PbZrO3 and PbTiO3/SrTiO3 (001) heterostructures.
The present paper is structured in the following way. Our hybrid-DFT calculations details, dealing with ABO3 perovskite (001) surfaces and heterostructures, are described in Section 2. Section 3.1 presents our computation results dealing with SrZrO3, PbZrO3, PbTiO3, BaTiO3 and SrTiO3 (001) surfaces, whereas Section 3.2 addresses their SrTiO3/BaTiO3, SrZrO3/PbZrO3 and PbTiO3/SrTiO3 (001) heterostructures. In particular, our novel hybrid-DFT computations are devoted to charge density redistribution and changes in the band structure in stoichiometric, as well as non-stoichiometric (001), heterostructures, consisting of BTO, PTO and PZO (001) thin films, with different thicknesses deposited on the TiO2-, TiO2- and ZrO2-terminated STO and SZO (001) substrate. We analyzed our hybrid-DFT calculation results and detected systematic trends common for all STO/BTO and STO/PTO, as well as SZO/PZO (001), heterostructures.

2. Hybrid-DFT Calculation Details

We carried out our hybrid-DFT computations for the PbTiO3 (PTO), BaTiO3 (BTO), SrTiO3 (STO), PbZrO3 (PZO) and SrZrO3 (001) bulk, as well as their BO2- and AO-terminated (001) surfaces, and their different (001) heterostructures, employing the computer code CRYSTAL [97]. For our hybrid-DFT computations, we employed the nowadays very popular hybrid exchange-correlation functionals B3LYP [98] or B3PW [99,100]. It is worthwhile to note that the hybrid exchange-correlation functionals, for example, B3LYP or B3PW allows achieving an outstanding agreement with the experiment for the band gaps of ABO3 perovskites as well as related materials, like CaF2 [101] and MgF2 [102]. In contrast, density functional theory (DFT) strongly underestimates, while the Hartree–Fock (HF) method, as a rule, very strongly overestimates the ab initio calculated band gaps of solids. For example, the experimental CaF2 direct band gap value at the Γ-point is equal to 12.1 eV [103]. Our, by means of Perdew–Burke–Ernzerhof (PBE) exchange-correlation functional [104,105], extensively used in DFT, ab initio computed CaF2 Γ-Γ band gap is very strongly underestimated with respect to the experimental value (12.1 eV) and is equal to 8.45 eV. From another side, our ab initio HF [106] computed CaF2 Γ-Γ band gap (20.77 eV) is 1.72 times overestimated compared to the experimental CaF2 band gap value (12.1 eV [105]). According to our hybrid-DFT computations, the hybrid exchange correlation functionals, like B3LYP (10.85 eV) and B3PW (10.96 eV) allows the achievement of the best possible agreement with the experimental CaF2 Γ-Γ band gap equal to 12.1 eV. The experimentally measured MgF2 Γ-Γ band gap is equal to 13.0 eV [107]. Again, our PBE exchange-correlation functional computed MgF2 Γ-Γ band gap is considerably underestimated regarding the experimental value of (13.0 eV) and is equal to 6.9 eV. In contrast, our ab initio HF computed MgF2 Γ-Γ band gap (19.65 eV) very strongly, namely 1.65 times, overestimates the experimentally measured MgF2 band gap at Γ-point (13.0 eV). The best possible results, according to our hybrid-DFT computations, are possible to achieve by means of the hybrid exchange-correlation functionals B3LYP (9.42 eV) and B3PW (9.48 eV). Since, as it is possible to see from our computation results for CaF2 and MgF2 solids, the hybrid exchange-correlation functionals B3LYP and B3PW allows achieving the best possible results for CaF2 and MgF2 Γ-Γ band gaps, we performed all of our coming ABO3 perovskite (001) surface and interface computations using the B3LYP or B3PW functionals. The B3LYP and B3PW hybrid exchange-correlation functionals make use of 20% of the Hartree–Fock method as well as 80% of the density functional Hamiltonian, when incorporated in the CRYSTAL computer code [97].
The major advantage of the CRYSTAL computer software package, which is of key importance for our hybrid-DFT computations of neutral SrZrO3, PbZrO3, PbTiO3, BaTiO3 and SrTiO3 (001) surfaces and interfaces, is the implementation of the isolated 2-dimensional (2D) slab model. In our hybrid-DFT computations, the reciprocal space integration was carried out by examining the Brillouin zone with the 8 × 8 × 8 times expanded Pack–Monkhorst net for the bulk of SrZrO3, PbZrO3, PbTiO3, BaTiO3 and SrTiO3 solids, as well as by 8 × 8 × 1 net for their (001) surfaces and interfaces. With a goal of reaching a high accuracy in our ab initio computations, we employed sufficiently large tolerances equal to 7, 8, 7, 7, 14 for the Coulomb overlap, Coulomb penetration and exchange overlap, as well as first exchange pseudo-overlap and the second exchange pseudo-overlap [97].
With the aim of computing the neutral BO2-terminated (001) surfaces of ABO3 solids, we customized symmetrical slabs. They involved nine neutral as well as alternating BO2 and AO layers (Figure 1). All of these nine layers were perpendicular to the [001] crystal direction [108,109,110,111]. Taking into consideration the standard ionic charges equal to (+2e) for the A atom and (+4e) for the B atom, as well as (−2e) for the O atom, both alternating BO2 and AO layers are neutral, since they have a summary slab charge equal to zero. The slab, containing nine layers, used by us for the BO2-terminated ABO3 perovskite (001) surface computations (Figure 1) consisted of a supercell containing 23 atoms. Our hybrid-DFT computed BO2-terminated (001) slabs (Figure 1) of ABO3 perovskite were non-stoichiometric. They have a unit cell, used in our hybrid-DFT computations, described by the following chemical formula—A4B5O14 [112,113,114].
The second slab in our hybdid-DFT computations was terminated from both sides by the AO planes (Figure 2). It also contained nine alternating layers (Figure 2) but consisted of a supercell that incorporated only 22 atoms. Furthermore, the AO-terminated nine-layer slab, used in our hybrid-DFT computations, was non-stoichiometric and has the following chemical formula—A5B4O13. For Sr, Ba, Pb, Ti and O atoms [115], we used the basis sets developed in [117]. The inner-core electrons for Sr, Ba, Pb, Ti and Zr atoms were described by a small core Hay–Wadt effective pseudopotentials [97,116]. We computed the number of light oxygen atoms using the all-electron basis sets [97,115]. In order to correctly define the chemical bonding in ABO3 perovskites, as well as covalency effects, we employed a widely accepted Mulliken population [117,118,119] analysis. The Mulliken population analysis is incorporated in the CRYSTAL computer code [97], used in all of our hybrid-DFT computations for the ABO3 solid (001) surfaces and their respective heterostructures (Figure 3).
With ambition to compute the ABO3 solid (001) surface energies, we opened our hybid-DFT computations with cleavage energies for unrelaxed AO- as well as BO2-terminated (001) surfaces. In our hybrid-DFT computations, the two nine-layer AO- as well as BO2-terminated slabs contained 22 and 23 atoms of each of them. These two AO- and BO2-terminated slabs represented together nine ABO3 perovskite bulk unit cells. Each ABO3 perovskite bulk unit cell contained five atoms. AO as well as BO2-terminated ABO3 perovskite (001) surfaces arose at the same time moment under crystal cleavage. Therefore, the relevant cleavage energy is shared uniformly among created surfaces. Thereby, our ab initio computed ABO3 perovskite (001) surface cleavage energy is equal for both AO- as well as BO2-terminations:
Esurfunrel = ¼ [Eslabunrel (AO) + Eslabunrel (BO2) − 9 Ebulk]
where Eslabunrel (AO) and Eslabunrel (BO2) are our hybrid-DFT computed total energies for unrelaxed AO- and BO2-terminated ABO3 perovskite nine-layer slabs. Ebulk is our hybrid-DFT computed total bulk unit cell energy containing five atoms. Factor 4 in Equation (1) arises from the fact that we created four surfaces during the ABO3 perovskite cleavage event. As a next step, we computed the relaxation energies for AO- and BO2-terminated ABO3 solid (001) surfaces. We relaxed both sides of our nine-layer AO- and BO2-terminated slabs.
Erel(Ψ) = ½ [Eslab (Ψ) − Eslabunrel (Ψ)]
Eslab(Ψ) is our hybrid-DFT computed slab total energy after the geometry relaxation. In our case the symbol Ψ denotes AO- or BO2-terminated ABO3 perovskite (001) surface. After all of our hybrid-DFT computed ABO3 perovskite AO- or BO2-terminated (001) surface energy is defined as a sum of the cleavage as well as relaxation energies:
Esurf(Ψ) = Esurfunrel + Erel(Ψ)
As a next step, we will discuss our calculation details for the ABO3 perovskite (001) heterostructures, using as an example the BaTiO3/SrTiO3 (001) interface. At room temperature, the SrTiO3 substrate has a high-symmetry cubic structure. In our hybrid-DFT computations, we computed both SrTiO3 and BaTiO3 perovskites at their cubic, high-symmetry phase with the space group number P m 3 ¯ m . In our hybrid-DFT computations, we modelled the BaTiO3/SrTiO3 (001) interface employing the single-slab model. In order to maximally apply the advantages of symmetry, our slabs were symmetrically terminated. In our hybrid-DFT computations, the SrTiO3 (001) substrate always contained 11 alternating SrO as well as TiO2 layers. From 1 to 10 BaO and TiO2 alternating layers were augmented on both sides of the 11-layer TiO2-terminated SrTiO3 (001) substrate (Figure 3).
We allowed all coordinated of atoms to relax in our hybrid-DFT computed BaTiO3/SrTiO3 (001) heterostructure. In our hybrid-DFT computations, due to restrictions imposed by the cubic symmetry of the system, atomic displacements are possible only along axis z. It is worthwhile to note that the mismatch equal to approximately 2.5 percent among BaTiO3 and SrTiO3 bulk lattice constants happens at the time of BaTiO3 epitaxial growth. The joint equilibrium average lattice constant employed in all of our future hybrid-DFT computations for BaTiO3/SrTiO3 (001) interfaces is equal to 3.958 Å. The joint lattice constant for the BaTiO3/SrTiO3 (001) interface was hybrid-DFT computed by us for the thickest (001) interface, which consisted of the 11-layer SrTiO3 substrate as well as, from both sides of this substrate, augmented 10 BaTiO3 layers. In our hybrid-DFT calculations of the shift (Δz) for each layer of the BaTiO3/SrTiO3 (001) heterostructure, we take into consideration the shift of the previous atomic layer. Therefore, the reference z coordinate for each mono-layer N is described by the following equation:
zNref = ½ [zN-1Me + zN-1O)
where, in Equation (4), zN-1Me as well as zN-1O defines the z coordinates for the cation and anion from the previous atomic monolayer.

3. Hybrid-DFT Computation Results

3.1. Pristine SrTiO3, BaTiO3, PbTiO3, SrZrO3 and PbZrO3 (001) Surfaces

As an opening to our hybrid-DFT computations, we computed the bulk lattice constants for five ABO3 perovskites. Our hybrid-DFT computed bulk lattice constants for PbTiO3 (3.936 Å), BaTiO3 (4.008 Å), SrTiO3 (3.904 Å), PbZrO3 (4.220 Å) and SrZrO3 (4.195 Å) are in fair agreement with the experimentally detected bulk lattice constants for PbTiO3 (3.97 Å [120]), BaTiO3 (4.004 Å [121]), SrTiO3 (3.898 Å [122]), PbZrO3 (4.1614 Å [123]) and SrZrO3 (4.154 Å [21]). Our hybrid-DFT computed effective atomic charges, as well as chemical bond populations for the PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 perovskite bulk, are collected in Table 1.
As we can see from Table 1, our hybrid-DFT computed atomic charges for all five perovskites are considerably smaller than those expected in an ionic model for the A atom (+2e), for the B atom (+4e) and also for the O atom (−2e). For example, our hybrid-DFT computed effective atomic charges for the Pb atom, in the different compounds PbTiO3 (+1.354e) and PbZrO3 (+1.368e), are almost equal. The same is also true for the Sr atom. Furthermore, the Sr atom atomic charges in different perovskites, such as SrTiO3 (+1.871e) and SrZrO3 (+1.880e), are almost equal (Table 1). The Zr atom effective charges in the PbZrO3 (+2.111e) and SrZrO3 (+2.174e) perovskites are almost equal and are always considerably smaller than the Ti atom effective charges in the PbTiO3 (+2.341e), BaTiO3 (+2.367e) and SrTiO3 (+2.351e) crystals. It is worthwhile to note that the Ti atom effective charges in PbTiO3, BaTiO3 and SrTiO3 perovskites (+2.341e, +2.367e and +2.351e, respectively) almost coincide (Table 1). The chemical bond covalency between the B and O atoms are in the range from (+0.088e) between Ti-O atoms in the SrTiO3 perovskite to (+0.106e) between Zr-O atoms in the PbZrO3 crystal. The bond populations between Ti-O atoms are equal to (0.098e) in both PbTiO3 and BaTiO3 perovskites.
As a next step, we computed the bulk band gaps at the Γ-point for PbTiO3, BaTiO3 and SrTiO3 perovskites and plotted the band structure for BaTiO3 and PbTiO3 crystals. We compared our hybrid-DFT computation results for Γ-Γ band gaps with available experimental results. Our B3PW computed SrTiO3 bulk Γ-Γ band gap is equal to 3.96 eV, which is in fair agreement with the available experimental data for SrTiO3 direct Γ-Γ bulk band gap equal to 3.75 eV [124] (Table 2). The direct Γ-Γ bulk band gap in the BaTiO3 perovskite is measured experimentally (Table 2) at the tetragonal to orthorhombic phase transition temperature equal to 278 K. The BaTiO3 direct bulk Γ-Γ band gap in different experimental conditions is slightly different and is equal to 3.27 eV or 3.38 eV, respectively [125]. Our B3PW computed BaTiO3 direct bulk Γ-Γ band gap (3.55 eV) (Table 2) is in almost perfect agreement with these experiments (Figure 4a). Finally, our B3PW computed PbTiO3 bulk band gap at Γ-point (4.32 eV) (Figure 4b) (Table 2) is in satisfactory agreement with the experimentally detected direct PbTiO3 bulk Γ-Γ band gap (3.4 eV) [126] at the PbTiO3 perovskite tetragonal phase.
As we can see from Table 3, according to our hybrid-DFT computations, all upper-layer atoms for all five perovskites relax inwards in the direction towards the perovskite bulk. The only exception from this systematic trend is the upward relaxation of the TiO2-terminated PbTiO3 (001) surface upper-layer O atom by 0.31 % of a0. Just opposite, all second-layer atoms relax upwards. For both upper layers, the metal atom relaxation magnitudes are always larger than the respective oxygen atom relaxation magnitude [127,128,129,130]. The only exception to this systematic trend is the ZrO2-terminated SrZrO3 (001) surface upper-layer oxygen atom relaxation magnitude (−2.10% of a0), which is slightly larger than the same-layer Sr atom relaxation magnitude (−1.38% of a0) (Table 3).
In most cases, as it is possible to see from our hybrid-DFT computation results collected in Table 4 for AO-terminated PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 (001) surfaces, all upper-layer atoms relax inwards (Table 4), whereas all second-layer atoms relax upwards. The only two exceptions from this systematic trend are the upward relaxation of O atom on the upper layer of SrO-terminated SrTiO3 (001) surface by (+0.84% of a0), as well as the inward relaxation of the second-layer O atom on the SrO-terminated SrZrO3 (001) surface by a very small relaxation magnitude of only (−0.05 of a0) (Table 4). It is interesting to notice that for the AO-terminated PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 (001) surfaces, the relaxation magnitudes of metal atoms in the upper, as well as the second, surface layers are always larger than the relaxation magnitudes of the respective oxygen atoms (Table 4).
As we can see from Table 5, our hybrid-DFT computed surface rumpling s (Figure 1) is always positive for BO2-terminated ABO3 perovskite (001) surfaces. The definitions of surface rumpling s, as well as the near-surface interplanar separations Δd12 and Δd23, are explained in Figure 1. The only exception to this systematic trend is our hybrid-DFT computed negative surface rumpling s for the ZrO2-terminated SrZrO3 (001) surface (−0.72) (Table 5). Moreover, the surface rumpling s computed by Wang et al. [131] for the ZrO2-terminated SrZrO3 (001) surface is negative (Table 5) (−0.70) and almost coincides with our result (−0.72). The experimental data for the surface rumpling s are available, for example, for the SrTiO3 crystal. Our B3PW computed surface rumpling s for SrTiO3 perovskite (+2.12) is in fair agreement with the available LEED [132] (+2.1 ± 2), as well as RHEED [133] (+2.6), experiments (Table 5). It is worthwhile to note that all of our hybrid-DFT computed interlayer distances for all materials (Figure 1) Δd12 are negative. This means that, according to our hybrid-DFT computations, the contraction always happens between the first and second surface layer for all BO2-terminated ABO3 perovskite (001) surfaces (Table 5). Oppositely, according to our hybrid-DFT computations, there is always expansion between the second and third surface layer Δd23 (Figure 1) for all of our hybrid-DFT computed BO2-terminated ABO3 perovskite (001) surfaces (Table 5).
As we can see from our hybrid-DFT computation results, collected in Table 6, the B-O atom chemical bond populations on the BO2-terminated PbZrO3, SrZrO3, PbTiO3, BaTiO3 and SrTiO3 perovskite (001) surfaces (+0.116e, +0.114e, +0.114e, +0.126e, +0.118e, respectively) are always larger than the B-O chemical bond population at the same perovskite bulk (+0.106e, +0.092e, +0.098e, +0.098e, +0.088e, respectively). As we can see from Table 6, the largest chemical bond population is between the Ti-O atoms (+0.126e) on the TiO2-terminated BaTiO3 (001) surface, whereas the smallest chemical bond population (+0.114e) is between the Zr-O and Ti-O atoms on the SrZrO3, as well as the BaTiO3, ZrO2 and TiO2-terminated (001), surfaces. It is worthwhile to note that the Ti atom’s effective charges on the TiO2-terminated PbTiO3, BaTiO3 and SrTiO3 (001) surfaces (+2.279e, +2.307e and +2.291e, respectively) are reduced with regard to the Ti atom’s effective charges in the PbTiO3, BaTiO3 and SrTiO3 bulk (+2.341e, +2.367e and +2.351e, respectively). Just the opposite situation occurs for the PbZrO3 and SrZrO3 perovskites, wherein the Zr atom’s effective charge on the ZrO2-terminated PbZrO3 and SrZrO3 (001) surfaces (+2.165e and +2.196e, respectively) is larger than in the PbZrO3 and SrZrO3 bulk matrixes (+2.111e and +2.174e, respectively) (Table 6).
Our hybrid-DFT computed (001) surface cleavage and relaxation, as well as surface energies, for PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 perovskites are collected in Table 7. As we can see from Table 7, the largest cleavage energy is for the SrZrO3 perovskite (001) surface (1.56 eV), whereas the smallest cleavage energy is for the PbTiO3 perovskite (001) surface (1.02 eV) (Table 7). The largest surface relaxation energy is for the SrO-terminated SrZrO3 perovskite (001) surface (−0.43), but the smallest surface relaxation energy is for the BaO-terminated BaTiO3 (001) surface and is only (−0.11 eV) (Table 7). Finally, our hybrid-DTT computed surface energy is largest for the ZrO2-terminated SrZrO3 (001) surface (1.24 eV). Oppositely, our hybrid-DFT computed surface energy is the smallest for the TiO2-terminated PbTiO3 (001) surface and is only (0.74 eV). The largest energy difference (0.12 eV) for single perovskite (001) surfaces with different terminations is between the TiO2-terminated BaTiO3 (001) surface (1.07 eV) and the BaO-terminated BaTiO3 (001) surface (1.19 eV). This means, that, for all of our hybrid-DFT computed PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 (001) surfaces, the surface energies for one perovskite’s both surface terminations are rather close, and thereby both BO2 and AO (001) surface terminations may exist simultaneously for all our hybrid-DFT computed ABO3 perovskites (Table 7).
Our B3PW computed bulk band gaps (Figure 4) for PbTiO3, BaTiO3 and SrTiO3 perovskite bulk (4.32 eV, 3.55 eV and 3.96 eV, respectively) are in fair agreement with the available experimental data [124,125,134] (Table 8). As we can see from Table 8 and Figure 5 and Figure 6, our B3PW computed (001) surface band gaps for all three perovskites and both (001) surface terminations are always reduced with respect to the PbTiO3, BaTiO3 and SrTiO3 bulk. Namely, our hybrid-DFT computed TiO2-terminated PbTiO3, BaTiO3 and SrTiO3 (001) surface band gaps are equal to 3.18 eV, 2.96 eV (Figure 5) and 3.95 eV, respectively. At the same time, our hybrid-DFT computed AO-terminated PbTiO3, BaTiO3 and SrTiO3 (001) surface band gaps are equal to 3.58 eV, 3.49 eV (Figure 6) and 3.72 eV, respectively, and are also smaller than the respective bulk band gap values at Γ-point (Table 8).

3.2. BaTiO3/SrTiO3, PbTiO3/SrTiO3 and SrZrO3/PbZrO3 (001) Interfaces

In order to start our hybrid-DFT computations for BTO/STO, PTO/STO and SZO/PZO (001) heterostructures, we computed their joint lattice constants. Our hybrid-DFT computed joint lattice constant for BTO/STO (001) interface is equal to 3.958 Å. This joint BTO/STO (001) interface joint lattice constant was computed for the system containing an 11-layer thick STO substrate, as well as 10 BTO monolayers augmented on both sides of the STO substrate (Figure 3). Thereby, our B3PW computed thickest BTO/STO (001) interface contained 31 monolayers, as well as 78 atoms (Figure 3) [135]. Using a similar model for the PTO/STO (001) thickest interface, our B3PW computed joint lattice constant for this system is equal to 3.91 Å. Finally, our hybrid-DFT computed SZO/PZO (001) interface joint lattice constant for the system containing an 11-layer thick SZO substrate, as well as 10 PZO monolayers augmented symmetrically on both sides of the substrate, is equal to 4.167 Å.
In our hybrid-DFT (001) heterostructure computations, we relaxed all atomic positions only alongside the z-axis, due to the symmetry constraints for the cubic ABO3 perovskite matrixes. We computed the atomic shifts Δz regarding the averaged coordinate z of the former atomic layer as described in Equation (4). As we can see from Figure 7, our B3PW computed upper SrTiO3 substrate (001) layer, which contains Ti and O atoms (x = 0) in (Figure 7), relaxes very strongly inside (−5.95% of a0). According to our B3PW computations, upper BTO (001) layer atoms, augmented on the STO 11-layer (001) substrate, always very strongly shifts (Δz) inwards (x = 1–10 in Figure 7). It is worthwhile to note that the BTO upper-layer atoms inwards-relaxation numerical value Δz considerably depends on the number of the STO (001) 11-layer substrate augmented BTO (001) layers.
In the case of one augmented BTO (001) layer, consisting of the BaO atoms, the atom relaxation magnitude Δz is equal to −1.54% of the joint lattice constant a0 (Figure 3 and Figure 7). For two augmented BTO (001) layers, where the upper layer contains TiO2 atoms, the upper-layer atom relaxation magnitude Δz is equal to (−3.20% of a0) (Figure 7). For three augmented BTO (001) layers, the upper BaO layer atom relaxation magnitude Δz is equal to (−1.84% of a0), for 4 layers (−3.55% of a0), for five layers (−2.07% of a0), for six layers (−3.70% of a0), for seven layers (−2.21% of a0), for eight layers (−3.84% of a0), for nine layers (−2.32% of a0) and finally, for ten augmented BTO (001) layers, the upper TiO2 layer atom relaxation magnitude Δz is equal to (−3.92% of a0) (Figure 7).
As we can see from Figure 8, for one augmented PTO (001) layer (x = 1), consisting of the PbO atoms, on the 11-layer STO (001) substrate (x = 0 for the substrate upper layer TiO2 atoms relaxation magnitude Δz), the upper layer PbO atoms’ relaxation magnitude Δz is equal to (−6.01% of a0). For two augmented PTO (001) layers, the upper layer consists of TiO2 atoms (x = 2), and their relaxation magnitude Δz is equal to (−7.76% of a0). For three augmented PTO (001) layers (x = 3), the upper layer consists of PbO atoms, and their relaxation magnitude Δz is equal to (−6.97% of a0). As we can see from Figure 8, for four augmented PTO (001) layers (x = 4), the upper TiO2 atoms containing layer atoms’ relaxation magnitude Δz is equal to (−8.25% of a0). For (x = 5), the upper augmented PTO (001) layer atom relaxation magnitude Δz is equal to (−7.26% of a0), for (x = 6) (−8.38% of a0), for (x = 7) (−7.34% of a0), for (x = 8) (−8.54% of a0), for (x = 9) (−7.42% of a0) and for (x = 10) (−8.54% of a0) (Figure 8).
As we can see from Figure 9, our hybrid-DFT computed Γ-Γ band gap for the PTO/STO (001) interface, containing an 11-layer STO (001) substrate as well as a PbO (001) layer (x = 1), augmented on both sides of this substrate, is equivalent to 3.45 eV. For two PTO (001) layers, augmented from both sides on the 11-layer STO (001) substrate, the whole PTO/STO (001) interface system is terminated by the TiO2 (001) layers from both sides and has a Γ-Γ band gap equal to 3.18 eV (Figure 9). For three PTO (001) layers, augmented on the STO (001) substrate, the whole PTO/STO (001) interface is terminated by the PbO-terminated (001) layer from both sides and has a Γ-Γ band gap equal to 3.25 eV. As we can see from Figure 9, for four augmented PTO (001) layers (x = 4), the Γ-Γ band gap of PTO/STO (001) interface is equal to 3.17 eV, for 5 augmented layers (3.08 eV), for 6 layers (3.05 eV), for 7 layers (2.99 eV), for 8 layers (2.99 eV), for 9 layers (2.94 eV) and, finally, for 10 augmented PTO (001) layers, the Γ-Γ band gap of PTO/STO (001) interface is equal to (2.93 eV).
As we can see from Figure 10a,b, our hybrid-DFT computed Γ-Γ band gap for BTO/STO, as well as SZO/PZO (001) interface Γ-Γ band gap, very strongly depends on the upper augmented layer BO2- or AO-termination but considerably less so on the number of augmented (001) layers.

4. Conclusions

According to our hybrid-DFT computations for BO2- and AO-terminated PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 (001) surfaces, in most cases, the upper-layer ions relax inwards, towards the bulk, whereas the second-layer ions shift upwards. It is worthwhile to also note that independent calculation results of other authors for ABO3 perovskite material’s neutral (001) surfaces confirm this systematic trend, that the upper layer ions, in most cases, relax inwards, whereas the second layer ions, again, in most cases, relax upwards [4,9,20,49,50,54,66,67,135,136,137,138].
Our hybrid-DFT computed surface rumpling s for the BO2-terminated ABO3 perovskite (001) surfaces (Table 5) is positive for most computed ABO3 perovskites, such as PbZrO3, PbTiO3, BaTiO3 and SrTiO3 (+0.38, +3.12, +2.73 and +2.12, respectively). Furthermore, the experimentally detected surface rumpling s for the TiO2-terminated SrTiO3 (001) surface matrix by two different experimental methods (LEED [132] and RHEED [133]) is positive (Table 5). Moreover, the agreement between our B3PW computed surface rumpling s for the TiO2-terminated SrTiO3 (001) surface (+2.12) and LEED [131] experimentally detected respective SrTiO3 (001) surface rumpling s (+2.1 ± 2) is almost perfect (Table 5). Nevertheless, there is a single exception to this systematic trend. Namely, our hybrid-DFT computed surface rumpling s for the ZrO2-terminated SrZrO3 (001) surface is negative (−0.72) and is in line with another ab initio calculation result performed by Wang et al. [131] (−0.70).
Our hybrid-DFT computed B-O atom chemical bond population values in the bulk of PbTiO3 (+0.098e), BaTiO3 (+0.098e), SrTiO3 (+0.088e), PbZrO3 (+0.106e) and SrZrO3 (+0.092e) perovskites are considerably enhanced on their BO2-terminated PbTiO3 (+0.114e), BaTiO3 (+0.126e), SrTiO3 (+0.118e), PbZrO3 (+0.116e) and SrZrO3 (+0.114e) (001) surfaces (Table 6). Our hybrid-DFT computed BO2- and AO-terminated (001) surface energies for PbTiO3 (0.74 and 0.83 eV), BaTiO3 (1.07 and 1.19 eV), SrTiO3 (1.23 and 1.15 eV), PbZrO3 (0.93 and 1.00 eV) and SrZrO3 (1.24 and 1.13 eV) are comparable for both (001) surface terminations (Table 7). Thereby, both BO2- and AO-terminated (001) surfaces may co-exist in ABO3 perovskite matrices. It is worthwhile to note that our hybrid-DFT computed ABO3 perovskite (001) surface energies are always smaller than the (011) [1,2] and especially the (111) [128] surface energies. This means, that, according to our hybrid-DFT computations, the ABO3 perovskite (001) surfaces are always the most stable, whereas the (011) and especially the (111) ABO3 perovskite surfaces are considerably less stable. Finally, our hybrid-DFT computed bulk Γ-Γ band gaps for PbTiO3, BaTiO3 and SrTiO3 perovskites are in fair agreement with the experimentally available data. For BO2- and AO-terminated PbTiO3, BaTiO3 and SrTiO3 (001) surfaces (Table 8), the band gaps are always reduced with regard to the respective bulk Γ-Γ band gaps.
For our B3PW computed BTO/STO as well as PTO/STO (001) interfaces, the average augmented upper-layer atom relaxation magnitudes increased by the number of augmented BTO or PTO (001) layers but always independently from the number of augmented layers, which were stronger for TiO2-terminated than BaO- or PbO-terminated upper augmented layers. All of our B3PW computed PTO/STO (001) interface upper augmented layer average atom displacement magnitudes Δz are between (−6.01% of a0) for the first augmented layer and (−8.54% of a0) for 10 augmented layers. In contrast to the PTO/STO (001) interfaces, for our hybrid-DFT computed BTO/STO (001) interfaces, the upper augmented layer average atom displacement magnitudes Δz are considerably smaller, and they are in the range between the (−1.54% of a0 (1 layer) to −3.92% of a0 (10 layers), respectively). Our B3PW computed BTO/STO, as well as SZO/PZO (001) interface Γ-Γ band gaps, very strongly depends on the upper augmented layer BO2- or AO-termination but considerably less so on the number of augmented (001) layers [135,139].
Summing up, all of our hybrid-DFT computed BTO/STO and PTO/STO, as well as PZO/SZO (001) heterostructures, are semiconducting. In general agreement with available experimental data [84], according to our hybrid-DFT computations, the (001) interface layer does not considerably influence the electronic structure of our studied heterostructures. At the same time, the termination of the deposited BTO and PTO, as well as PZO (001), thin films atop STO or SZO (001) substrates, respectively, may shift the band edges regarding the vacuum level and thereby reduce the (001) heterostructure band gap [84].

Author Contributions

Conceptualization, R.J. and A.I.P.; methodology, R.I.E.; software, S.P.; validation, R.I.E., R.J. and D.B.; formal analysis, A.I.P.; investigation, J.P.; resources, D.B.; data curation, S.P.; writing—original draft preparation, R.I.E., J.P., D.B., S.P., R.J. and A.I.P.; writing—review and editing, R.I.E., J.P., A.I.P., R.J., S.P. and D.B; visualization, S.P.; funding acquisition, J.P. All authors have read and agreed to the published version of the manuscript.

Funding

We acknowledge the financial support from the funder—Latvian Council of Science. The funding number is: Grant No. LZP-2020/1-0345. The Institute of Solid State Physics, University of Latvia (Latvia), as the Centre of Excellence, has received funding from the European Unions Horizon 2020 Framework Programme H2020-WIDESPREAD01-2016-2017-Teaming Phase2 under Grant Agreement No. 739508, project CAMART-2.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Eglitis, R.I.; Vanderbilt, D. Ab initio calculations of the atomic and electronic structure of CaTiO3 (001) and (011) surfaces. Phys. Rev. B 2008, 78, 155420. [Google Scholar] [CrossRef] [Green Version]
  2. Eglitis, R.I.; Vanderbilt, D. Ab initio calculations of BaTiO3 and PbTiO3 (001) and (011) surface structures. Phys. Rev. B 2007, 76, 155439. [Google Scholar] [CrossRef] [Green Version]
  3. Erdman, N.; Poeppelmeier, K.R.; Asta, M.; Warschkov, O.; Ellis, D.E.; Marks, L.D. The structure and chemistry of the TiO2-rich surface of SrTiO3 (001). Nature 2002, 419, 55–58. [Google Scholar] [CrossRef]
  4. Dawber, M.; Rabe, K.M.; Scott, J.F. Physics of thin-film ferroelectric oxides. Rev. Mod. Phys. 2005, 77, 1083–1130. [Google Scholar] [CrossRef] [Green Version]
  5. Piyanzina, I.I.; Pavlov, D.P.; Jagličić, Z.; Shulyaev, D.A.; Tayurskii, D.A.; Kabanov, V.V.; Mamin, R.F. Structural and magnetic properties of ferroelectric/dielectric BaTiO3/LaMnO3 and BaTiO3/SrTiO3 heterostructures. Ferroelectrics 2021, 575, 144–150. [Google Scholar] [CrossRef]
  6. Eglitis, R.I.; Kleperis, J.; Purans, J.; Popov, A.I.; Jia, R. Ab initio calculations of CaZrO3 (011) surfaces: Systematic trends in polar (011) surface calculations of ABO3 perovskites. J. Mater. Sci. 2020, 55, 203–217. [Google Scholar] [CrossRef]
  7. Danyushkina, L.A.; Khaliullina, A.S.; Kuimov, V.M.; Osinkin, D.A.; Antonov, B.D.; Pankratov, A.A. Influence of modification of chemical solution deposition on morphology and conductivity of CaZr0.9Y0.1O3−δ films. Solid State Ion. 2019, 329, 1–7. [Google Scholar] [CrossRef]
  8. Eglitis, R.I.; Popov, A.I. Ab initio calculations for the polar (001) surfaces of YAlO3. Nucl. Instr. Meth. B 2018, 434, 1–5. [Google Scholar] [CrossRef]
  9. Brik, M.G.; Ma, C.G.; Krasnenko, V. First-principles calculations of the structural and electronic properties of the cubic CaZrO3 (001) surfaces. Surf. Sci. 2013, 608, 146–153. [Google Scholar] [CrossRef]
  10. Eglitis, R.I. Comparative ab initio calculations of SrTiO3 and CaTiO3 polar (111) surfaces. Phys. Stat. Sol. B 2015, 252, 635–642. [Google Scholar] [CrossRef]
  11. Ahmadi, F.; Aragi, H. Coexistence of magnetism and ferroelectricity in (PbTiO3)m/(BaTiO3)n superlattices. Superlattices Microstruct. 2020, 140, 106427. [Google Scholar] [CrossRef]
  12. Jia, W.; Vikhnin, V.S.; Liu, H.; Kapphan, S.; Eglitis, R.; Usvyat, D. Critical effects in optical response due to charge transfer vibronic excitions and their structure in perovskite-like systems. J. Lumin. 1999, 83, 109–113. [Google Scholar] [CrossRef]
  13. Zhang, X.; Demkov, A.A.; Li, H.; Hu, X.; Wei, Y.; Kulik, J. Atomic and electronic structure of the Si/SrTiO3 interface. Phys. Rev. B 2003, 68, 125323. [Google Scholar] [CrossRef]
  14. Eglitis, R.I.; Purans, J.; Popov, A.I.; Bocharov, D.; Chekhovska, A.; Jia, R. Ab initio computations of O and AO as well as ReO2, WO2 and BO2-terminated ReO3, WO3, BaTiO3, SrTiO3 and BaZrO3 (001) surfaces. Symmetry 2022, 14, 1050. [Google Scholar] [CrossRef]
  15. Raza, S.; Zhang, R.; Zhang, N.; Li, Z.; Liu, L.; Zhang, F.; Wang, D.; Jia, C.L. ATiO3/TiO (A=Pb, Sn) superlattice: Bridging ferroelectricity and conductivity. Comput. Condens. Matter 2020, 25, e00491. [Google Scholar] [CrossRef]
  16. Borstel, G.; Eglitis, R.I.; Kotomin, E.A.; Heifets, E. Modelling of defects and surfaces in perovskite ferroelectrics. Phys. Stat. Sol. B 2003, 236, 253–264. [Google Scholar] [CrossRef]
  17. Cohen, R.E. Origin of ferroelectricity in perovskite oxides. Nature 1992, 358, 136–138. [Google Scholar] [CrossRef]
  18. Zhong, W.; Vanderbilt, D.; Rabe, K.M. Phase transitions in BaTiO3 from first principles. Phys. Rev. Lett. 1994, 73, 1861–1864. [Google Scholar] [CrossRef] [Green Version]
  19. Eglitis, R.I.; Kotomin, E.A.; Borstel, G. Quantum chemical modelling of perovskite solid solutions. J. Phys. Condens. Matter 2000, 12, L431–L434. [Google Scholar] [CrossRef] [Green Version]
  20. Sambrano, J.R.; Longo, V.M.; Longo, E.; Taft, C.A. Electronic and structural properties of the (001) SrZrO3 surface. J. Mol. Struct. THEOCHEM 2007, 813, 49–56. [Google Scholar] [CrossRef]
  21. Kennedy, B.J.; Howard, C.J. High-temperature phase transition in SrZrO3. Phys. Rev. B 1999, 59, 4023–4027. [Google Scholar] [CrossRef]
  22. Maniwa, H.; Sawaguchi, E.; Hoshino, S. Antiferroelectric structure of lead zirconate. Phys. Rev. 1951, 83, 1078. [Google Scholar]
  23. Shirane, G.; Sawaguchi, E.; Takagi, Y. Dielectric properties of lead zirconate. Phys. Rev. 1951, 84, 476–481. [Google Scholar] [CrossRef]
  24. Wang, H.; Liu, Y.; Yang, T.; Zhang, S. Ultra high energy-storage density in antiferroelectric ceramics with field-induced multiphase transitions. Adv. Funct. Mater. 2019, 29, 1807321. [Google Scholar] [CrossRef]
  25. Chen, M.J.; Ning, X.K.; Wang, S.F.; Fu, G.S. Significant enhancement of energy storage density and polarization in self-assembled PbZrO3: NiO nano-columnar composite films. Nanoscale 2019, 11, 1914–1920. [Google Scholar] [CrossRef]
  26. Zhang, S.; Li, F. High performance ferroelectric relaxor-PbTiO3 single crystals: Status and perspective. Appl. Phys. Rev. 2012, 111, 031301. [Google Scholar] [CrossRef] [Green Version]
  27. Sebastian, M.T. Polymer-Ceramic Composites of O-3 Connectivity for Circuits in Electronics: A Review. Int. J. Appl. Ceram. Technol. 2010, 7, 415–434. [Google Scholar] [CrossRef]
  28. Tan, G.; Wang, S.; Zhu, Y.; Zhou, L.; Yu, P.; Wang, X.; He, T.; Chen, J.; Mao, C.; Ning, C. Surface-selective preferential production of reactive oxygen species on piezoelectric ceramics for bacterial killing. ACS Appl. Mater. Interfaces 2016, 8, 24306–24309. [Google Scholar] [CrossRef] [Green Version]
  29. Kumar, S.; Vaish, R.; Powar, S. Surface-selective bactericidal effect on poled ferroelectric materials. J. Appl. Phys. 2018, 124, 014901. [Google Scholar] [CrossRef]
  30. Acosta, M.; Novak, N.; Rojas, V.; Patel, S.; Vaish, R.; Koruza, J.; Rossetti, G.A.; Rödel, J. BaTiO3-based piezoelectrics: Fundamentals, current status, and perspectives. Appl. Phys. Rev. 2017, 4, 041305. [Google Scholar] [CrossRef] [Green Version]
  31. Domen, K.; Kudo, A.; Onishi, T.; Kosugi, N.; Kuroda, H. Photocatalytic decomposition of water into hydrogen and oxygen over nickel (II) oxide-strontium titanate (SrTiO3) powder. 1. Structure of the catalysts. J. Phys. Chem. 1986, 90, 292–295. [Google Scholar] [CrossRef]
  32. Ahuja, S.; Kutty, T.R.N. Nanoparticles of SrTiO3 prepared by gel to crystallite conversion and their photocatalytic activity in the mineralization of phenol. J. Photochem. Photobiol. A Chem. 1996, 97, 99–107. [Google Scholar] [CrossRef]
  33. Samsonidze, G.; Kozinsky, B. Accelerated Screening of Thermoelectric Materials by First-Principles Computations of Electron-Phonon Scattering. Adv. Energy Mater. 2018, 8, 1800246. [Google Scholar] [CrossRef] [Green Version]
  34. Guo, R.; Wang, X.; Kuang, Y.; Huang, B. First-principles study of anisotropic thermoelectric transport properties of IV-VI semiconductor compounds SnSe and SnS. Phys. Rev. B 2015, 92, 115202. [Google Scholar] [CrossRef] [Green Version]
  35. Xi, L.; Zhang, Y.B.; Shi, Z.Y.; Yang, J.; Shi, X.; Chen, L.D.; Zhang, W.; Yang, J.; Sing, D.J. Chemical bonding, conductive network, and thermoelectric performance of the ternary semi conductors Cu2SnX3 (X = Se, S) from first principles. Phys. Rev. B 2012, 86, 155201. [Google Scholar] [CrossRef]
  36. Tian, Z.; Garg, J.; Esfarjani, K.; Shiga, T.; Shiomi, J.; Chen, G. Phonon conduction in PbSe, PbTe, and PbTe1−xSex from first principles calculations. Phys. Rev. B 2012, 85, 184303. [Google Scholar] [CrossRef] [Green Version]
  37. Orhan, E.; Pontes, F.M.; Pinteiro, C.D.; Boschi, T.M.; Leite, E.R.; Pizani, P.S.; Beltrán, A.; Andrés, J.; Varela, J.A.; Longo, E. Origin of photoluminescence in SrTiO3: A combined experimental and theoretical study. J. Solid State Chem. 2004, 177, 3879–3885. [Google Scholar] [CrossRef]
  38. Grigorjeva, L.; Millers, D.K.; Pankratov, V.; Williams, R.T.; Eglitis, R.I.; Kotomin, E.A.; Borstel, G. Experimental and theoretical studies of polaron optical properties in KNbO3 perovskite. Solid State Commun. 2004, 129, 691–696. [Google Scholar] [CrossRef]
  39. Devreese, J.T.; Fomin, V.M.; Pokatilov, E.P.; Kotomin, E.A.; Eglitis, R.; Zhukovskii, Y.F. Theory of bound polarons in oxide compounds. Phys. Rev. B 2001, 63, 184304. [Google Scholar] [CrossRef] [Green Version]
  40. Leonelli, R.; Brebner, J.L. Evidence for bimolecular recombination in the luminescence spectra of SrTiO3. Solid State Commun. 1985, 54, 505–507. [Google Scholar] [CrossRef]
  41. Makarova, M.V.; Prokhorov, A.; Stupak, A.; Kopeček, J.; Drahokoupil, J.; Trepakov, V.; Dejneka, A. Synthesis and Magnetic Properties of Carbon Doped and Reduced SrTiO3 Nanoparticles. Crystals 2022, 12, 1275. [Google Scholar] [CrossRef]
  42. Meng, J.; Huang, Y.; Zhang, W.; Du, Z.; Zhu, Z.; Zou, G. Photoluminescence in nanocrystalline BaTiO3 and SrTiO3. Phys. Lett. A 1995, 205, 72–76. [Google Scholar] [CrossRef]
  43. Rinnert, H.; Vergat, M.; Marchal, G.; Burneau, A. Intense visible photoluminescence in amorphous SiOx and SiOx:H films prepared by evaporation. Appl. Phys. Lett. 1998, 72, 3157. [Google Scholar] [CrossRef] [Green Version]
  44. Scott, J.F. Status report on ferroelectric memory materials. Integr. Ferroelectr. 1998, 20, 15–23. [Google Scholar] [CrossRef]
  45. Millis, A.J. Lattice effects in magnetoresistive manganese perovskites. Nature 1998, 392, 147–150. [Google Scholar] [CrossRef]
  46. Wessels, B.W. Metal-organic chemical vapor deposition of ferroelectric oxide thin films for electronic and optical applications. Annu. Rev. Mater. Sci. 1995, 25, 525–546. [Google Scholar] [CrossRef]
  47. Farlenkov, A.S.; Ananyev, M.V.; Eremin, V.A.; Porotnikova, N.M.; Kurumchin, E.K.; Melekh, B.T. Oxygen isotope exchange in doped calcium and barium zirconates. Solid State Ion. 2016, 290, 108–115. [Google Scholar] [CrossRef]
  48. Ananyev, M.V.; Bershitskaya, N.M.; Plaksin, S.V.; Kurumchin, E.K. Phase equilibriums, oxygen exchange kinetics and diffusion in oxides CaZr1−xScxO3−x/2−δ. Russ. J. Electrochem. 2012, 48, 879–886. [Google Scholar] [CrossRef]
  49. Celik, F.A. Electronic structure of two-dimensional-layered PbTiO3 perovskite crystal: An extended tight-binding study based on DFT. Bull. Mater. Sci. 2022, 45, 108. [Google Scholar] [CrossRef]
  50. Eglitis, R.I. Ab initio calculations of SrTiO3, BaTiO3, PbTiO3, CaTiO3, SrZrO3, PbZrO3 and BaZrO3 (001), (011) and (111) surfaces as well as F centers, polarons, KTN solid solutions and Nb impurities therein. Int. J. Mod. Phys. B 2014, 28, 1430009. [Google Scholar] [CrossRef] [Green Version]
  51. Noor, N.A.; Mahmood, Q.; Rashid, M.; Haq, B.V.; Laref, A.; Ahmad, S.A. Ab-initio study of thermodynamic stability, thermoelectric and optical properties of perovskites ATiO3 (A = Pb, Sn). J. Solid State Chem. 2018, 263, 115–122. [Google Scholar] [CrossRef]
  52. Shimada, T.; Tomoda, S.; Kitamura, T. Ab initio study of ferroelectricity in edged PbTiO3 nanowires under axial tension. Phys. Rev. B 2009, 79, 024102. [Google Scholar] [CrossRef] [Green Version]
  53. Zhang, J.M.; Pang, Q.; Xu, K.W.; Ji, V. First-principles study of the (110) polar surface of cubic PbTiO3. Comput. Mater. Sci. 2009, 44, 1360–1365. [Google Scholar] [CrossRef]
  54. Lazaro, S.D.; Longo, E.; Sambrano, J.R.; Beltrán, A. Structural and electronic properties of PbTiO3 slabs: A DFT periodic study. Surf. Sci. 2004, 552, 149–159. [Google Scholar] [CrossRef]
  55. Eglitis, R.I.; Purans, J.; Popov, A.I.; Jia, R. Tendencies in ABO3 perovskite and SrF2, BaF2 and CaF2 bulk and surface F-center ab initio computations at high symmetry cubic structure. Symmetry 2021, 13, 1920. [Google Scholar] [CrossRef]
  56. Gao, H.; Yue, Z.; Liu, Y.; Hu, J.; Li, X. A first-principles study on the multiferroic property of two-dimensional BaTiO3 (001) ultrathin film with surface Ba vacancy. Nanomaterials 2019, 9, 269. [Google Scholar] [CrossRef] [Green Version]
  57. Eglitis, R.I.; Purans, J.; Gabrusenoks, J.; Popov, A.I.; Jia, R. Comparative ab initio calculation of ReO3, SrZrO3, BaZrO3, PbZrO3 and CaZrO3 (001) surfaces. Crystals 2020, 10, 745. [Google Scholar] [CrossRef]
  58. Zhong, M.; Zeng, W.; Liu, F.S.; Tang, B.; Liu, Q.J. First-principles study of the atomic structures, electronic properties, and surface stability of BaTiO3 (001) and (011) surfaces. Surf. Interface Anal. 2019, 51, 1021–1032. [Google Scholar] [CrossRef]
  59. Eglitis, R.I.; Purans, J.; Jia, R. Comparative Hybrid Hartree-Fock-DFT Calculations of WO2-Terminated Cubic WO3 as well as SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces. Crystals 2021, 11, 455. [Google Scholar] [CrossRef]
  60. Costa-Amaral, R.; Gohda, Y. First-principles study of the adsorption of 3d transition metals on BaO- and TiO2-terminated cubic-phase BaTiO3 (001) surfaces. J. Chem. Phys. 2020, 152, 204701. [Google Scholar] [CrossRef]
  61. Sokolov, M.; Eglitis, R.I.; Piskunov, S.; Zhukovskii, Y.F. Ab initio hybrid DFT calculations of BaTiO3 bulk and BaO-terminated (001) surface F-centers. Int. J. Mod. Phys. B 2017, 31, 1750251. [Google Scholar] [CrossRef]
  62. Eglitis, R.I.; Piskunov, S. First principles calculations of SrZrO3 bulk and ZrO2-terminated (001) surface F centers. Comput. Condens. Matter 2016, 7, 1–6. [Google Scholar] [CrossRef] [Green Version]
  63. Charlton, G.; Brennan, S.; Muryn, C.A.; McGrath, R.; Norman, D.; Turner, T.S.; Thorton, G. Surface relaxation of SrTiO3 (001). Surf. Sci. 2000, 457, L376–L380. [Google Scholar] [CrossRef]
  64. Liang, Y.; Bonnell, D.A. Atomic structures of reduced SrTiO3 (001) surfaces. Surf. Sci. Lett. 1993, 285, L510–L516. [Google Scholar]
  65. Castell, M.R. Scanning tunneling microscopy of reconstructions on the SrTiO3 (001) surface. Surf. Sci. 2002, 505, 1–13. [Google Scholar] [CrossRef] [Green Version]
  66. Evarestov, R.A.; Bandura, A.V.; Alexandrov, V.E. Hybrid HF-DFT comparative study of SrZrO3 and SrTiO3 (001) surface properties. Phys. Stat. Sol. B 2006, 243, 2756–2763. [Google Scholar] [CrossRef]
  67. Iles, N.; Finocchi, F.; Khodja, K.D. A systematic study of ideal and double layered reconstructions of ABO3 (001) surfaces (A = Sr, Ba; B = Ti, Zr) from first principles. J. Phys. Condens. Matter 2010, 22, 305001. [Google Scholar] [CrossRef]
  68. Sopiha, K.V.; Malyi, O.I.; Persson, C.; Wu, P. Suppression of surface states at cubic perovskite (001) surfaces by CO2 adsorption. Phys. Chem. Chem. Phys. 2018, 20, 18828–18836. [Google Scholar] [CrossRef] [Green Version]
  69. Pilania, G.; Tan, D.Q.; Cao, Y.; Venkataramani, V.S.; Chen, Q.; Ramprasad, R. Ab initio study of antiferroelectric PbZrO3 (001) surfaces. J. Mater. Sci. 2009, 44, 5249–5255. [Google Scholar] [CrossRef]
  70. Ko, D.L.; Hsin, T.; Lai, Y.H.; Ho, S.Z.; Zheng, Y.; Huang, R.; Pan, H.; Chen, Y.C.; Chu, Y.H. High-stability transparent flexible energy storage based on PbZrO3 muscovite heterostructure. Nano Energy 2021, 87, 106149. [Google Scholar] [CrossRef]
  71. Guo, X.; Ge, J.; Ponchel, F.; Remiens, D.; Chen, Y.; Dong, X.; Wang, G. Effect of Sn substitution on the energy storage properties of high (001)-oriented PbZrO3 thin films. Thin Solid Film. 2017, 632, 93–96. [Google Scholar] [CrossRef]
  72. Qiao, L.; Song, C.; Sua, Y.; Fayaz, M.U.; Lu, T.; Yin, S.; Chen, C.; Xu, H.; Ren, T.L.; Pan, F. Observation of negative capacitance in antiferroelectric PbZrO3 Films. Nat. Commun. 2021, 12, 4215. [Google Scholar] [CrossRef] [PubMed]
  73. Delugas, P.; Filipetti, A.; Gadaleta, A.; Pallecchi, I.; Marre, D.; Fiorentini, V. Large band offset as a driving force of two-dimensional electron confinement: The case of SrTiO3/SrZrO3 interface. Phys. Rev. B 2013, 88, 115304. [Google Scholar] [CrossRef] [Green Version]
  74. Piskunov, S.; Eglitis, R.I. First principles hybrid DFT calculations of BaTiO3/SrTiO3 (001) interface. Solid State Ion. 2015, 274, 29–33. [Google Scholar] [CrossRef]
  75. Al-Aqtash, N.; Alsaad, A.; Sabirianov, R. Ferroelectric properties of BaZrO3/PbZrO3 and SrZrO3/PbZrO3 superlattices: An ab-initio study. J. Appl. Phys. 2014, 116, 074112. [Google Scholar] [CrossRef]
  76. Eglitis, R.; Kruchinin, S.P. Ab initio calculations of ABO3 perovskite (001), (011) and (111) nano-surfaces, interfaces and defects. Mod. Phys. Lett. B 2020, 34, 2040057. [Google Scholar] [CrossRef]
  77. Qi, H.; Chen, X.; Benckiser, E.; Wu, M.; Cristiani, G.; Logvenov, G.; Keimer, B.; Kaiser, V. Formation mechanism of Ruddlesden-Popper faults in compressive-strained ABO3 perovskite superlattices. Nanoscale 2021, 13, 20663–20669. [Google Scholar] [CrossRef] [PubMed]
  78. Aso, R.; Kan, D.; Shimakawa, Y.; Kurata, H. Atomic level observation of octahedral distortions at the perovskite oxide heterointerface. Sci. Rep. 2013, 3, 2214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Piskunov, S.; Eglitis, R.I. Comparative ab initio calculations of SrTiO3/BaTiO3 and SrZrO3/PbZrO3 (001) heterostructures. Nucl. Instr. Methods B 2016, 374, 20–23. [Google Scholar] [CrossRef]
  80. Fredrickson, K.D.; Demkov, A.A. Switchable conductivity at the ferroelectric interface: Nonpolar oxides. Phys. Rev. B 2015, 91, 115126. [Google Scholar] [CrossRef] [Green Version]
  81. Eglitis, R.I.; Piskunov, S.; Zhukovskii, Y.F. Ab initio calculations of PbTiO3/SrTiO3 (001) heterostructures. Phys. Stat. Sol. C 2016, 13, 913–920. [Google Scholar]
  82. Lebedev, A.I. Ground state and properties of ferroelectric superlattices based on crystals of the perovskite family. Phys. Solid State 2010, 52, 1448–1462. [Google Scholar] [CrossRef]
  83. Lebedev, A.I. Band offsets in heterojunctions formed by oxides with cubic perovskite structure. Phys. Solid State 2014, 56, 1039–1047. [Google Scholar] [CrossRef]
  84. Bi, Z.; Uberuaga, B.P.; Vernon, L.J.; Fu, E.; Wang, Y.; Li, N.; Wang, H.; Misra, A.; Jia, Q.X. Radiation damage in heteroepitaxial BaTiO3 thin films on SrTiO3 under Ne ion irradiation. J. Appl. Phys. 2013, 113, 023513. [Google Scholar] [CrossRef]
  85. Stepkova, V.; Marton, P.; Setter, N.; Hlinka, J. Closed-circuit domain quadruplets in BaTiO3 nanorods embedded in a SrTiO3 film. Phys. Rev. B 2014, 89, 060101. [Google Scholar] [CrossRef] [Green Version]
  86. Piyanzina, I.I.; Eyert, V.; Lysogorskiy, Y.V.; Tayurskii, D.A.; Kopp, T. Oxygen vacancies and hydrogen doping in LaAlO3/SrTiO3 heterostructures: Electronic properties and impact on surface and interface reconstruction. J. Phys. Condens. Matter 2019, 31, 295601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Piyanzina, I.I.; Pavlov, D.P.; Zagidullin, R.R.; Tayurskii, D.A.; Mamin, R.F. Structural and electronic properties of heterointerface composed of non-polar oxides: SrTiO3 and ferroelectric BaTiO3. Ferroelectrics 2019, 542, 7–12. [Google Scholar] [CrossRef]
  88. Piyanzina, I.I.; Lysogorskiy, Y.V.; Tayurskii, D.A.; Mamin, R.F. Electronic Properties of a Two-Dimensional Electron Gas at the Interface between Transition Metal Complex Oxides. Bull. Russ. Acad. Sci. Phys. 2018, 82, 234–237. [Google Scholar] [CrossRef]
  89. Nova, T.F.; Disa, A.S.; Fechner, M.; Cavalleri, A. Metastable ferroelectricity in optically strained SrTiO3. Science 2019, 364, 1075–1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Oliveira, M.C.; Ribeiro, R.A.P.; Longo, E.; Bomio, M.R.D.; Motta, F.V.; Lazaro, S.R.D. Temperature dependence on phase evolution in BaTiO3 polytypes studied using ab initio calculations. Int. J. Quantum Chem. 2020, 120, e26054. [Google Scholar] [CrossRef]
  91. Zelezny, V.; Chvostova, D.; Simek, D.; Maca, F.; Masek, J.; Setter, N.; Huang, Y.H. The variation of PbTiO3 bandgap at ferroelectric phase transition. J. Phys. Condens. Matter 2016, 28, 025501. [Google Scholar] [CrossRef] [PubMed]
  92. Ligny, D.D.; Richet, P. High-temperature heat capacity and thermal expansion of perovskites. Phys. Rev. B 1996, 53, 3013–3022. [Google Scholar] [CrossRef]
  93. Bungaro, C.; Rabe, K.M. Coexistence of antiferrodistortive and ferroelectric distortions at the BaTiO3 (001) surface. Phys. Rev. B 2005, 71, 035420. [Google Scholar] [CrossRef]
  94. Ghorbani, E.; Villa, L.; Erhart, P.; Klein, A.; Albe, K. Self-consistent calculations of charge self-trapping energies: A comparative study of polaron formation and migration in PbTiO3. Phys. Rev. Mater. 2022, 6, 074410. [Google Scholar] [CrossRef]
  95. Varley, J.B.; Janotti, A.; Franchini, C.; Walle, C.G.V. Role of self-trapping in luminescence and p-type conductivity of wide-band-gap oxides. Phys. Rev. B 2012, 85, 081109. [Google Scholar] [CrossRef] [Green Version]
  96. Eglitis, R.I.; Kotomin, E.A.; Trepakov, V.A.; Kapphan, S.E.; Borstel, G. Quantum chemical modelling of electron polarons and ‘green’ luminescence in PbTiO3 perovskite crystals. J. Phys. Condens. Matter 2002, 14, L647. [Google Scholar] [CrossRef] [Green Version]
  97. Dovesi, R.; Saunders, V.R.; Roetti, C.; Orlando, R.; Zicovich-Wilson, C.M.; Pascale, F.; Civalleri, B.; Doll, K.; Harrison, N.M.; Bush, I.J.; et al. CRYSTAL-2017 User Manual; University of Torino: Torino, Italy, 2017. [Google Scholar]
  98. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Perdew, J.P.; Wang, Y. Accurate and simple density functional for the electronic exchange energy: Generalized gradient approximation. Phys. Rev. B 1986, 33, 8800–8802, Erratum in Phys. Rev. B 1989, 40, 3399. [Google Scholar] [CrossRef] [PubMed]
  100. Perdew, J.P.; Wang, Y. Accurate and simple analytic representation of the electron-gas correlation energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef]
  101. Shi, H.; Eglitis, R.I.; Borstel, G. Ab initio calculations of the CaF2 electronic structure and F centers. Phys. Rev. B 2005, 72, 045109. [Google Scholar] [CrossRef]
  102. Vassilyeva, A.F.; Eglitis, R.I.; Kotomin, E.A.; Dauletbekova, A.K. Ab initio calculations of MgF2 (001) and (011) surface structure. Phys. B Condens. Matter 2010, 405, 2125–2127. [Google Scholar] [CrossRef]
  103. Rubloff, G.W. Far-Ultraviolet Reflectence Spectra and the Electronic Structure of Ionic Crystals. Phys. Rev. B 1972, 5, 662–684. [Google Scholar] [CrossRef]
  104. Perdew, J.; Ernzerhof, M.; Burke, K. Rationale for mixing exact exchange with density functional approximations. J. Chem. Phys. 1996, 105, 9982. [Google Scholar] [CrossRef]
  105. Perdew, J.P. Orbital functional for exchange and correlation: Self-interaction correction to the local density approximation. Chem. Phys. Lett. 1979, 64, 127–130. [Google Scholar] [CrossRef]
  106. Dovesi, R.; Orlando, R.; Roetti, C.; Pisani, C.; Saunders, V.R. The Periodic Hartree-Fock Method and Its Implementation in the Crystal Code. Phys. Stat. Sol. B 2000, 217, 63–88. [Google Scholar] [CrossRef]
  107. Lisitsyn, V.M.; Lisitsyna, L.A.; Popov, A.I.; Kotomin, E.A.; Abuova, F.U.; Akilbekov, A.; Maier, J. Stabilization of primary mobile radiation defects in MgF2 crystals. Nucl. Instr. Methods B 2016, 374, 24–28. [Google Scholar] [CrossRef]
  108. Heifets, E.; Kotomin, E.A.; Maier, J. Semi-empirical simulations of surface relaxation for perovskite titanates. Surf. Sci. 2000, 462, 19–35. [Google Scholar] [CrossRef]
  109. Heifets, E.; Eglitis, R.I.; Kotomin, E.A.; Maier, J.; Borstel, G. Ab initio modeling of surface structure for SrTiO3 perovskite crystals. Phys. Rev. B 2001, 64, 235417. [Google Scholar] [CrossRef] [Green Version]
  110. Ho, J.; Heifets, E.; Merinov, B. Ab initio simulation of the BaZrO3 (001) surface structure. Surf. Sci. 2007, 601, 490–497. [Google Scholar] [CrossRef]
  111. Heifets, E.; Dorfman, S.; Fuks, D.; Kotomin, E.; Gordon, A. Atomistic simulation of surface relaxation. J. Phys. Condens. Matter 1998, 10, L347. [Google Scholar] [CrossRef]
  112. Pilania, G.; Ramprasad, R. Adsorption of atomic oxygen on cubic PbTiO3 and LaMnO3 (001) surfaces. A density functional theory study. Surf. Sci. 2010, 604, 1889–1893. [Google Scholar] [CrossRef]
  113. Fechner, M.; Ostanin, S.; Mertig, I. Effect of the surface polarization in polar perovskites studied from first principles. Phys. Rev. B 2008, 77, 094112. [Google Scholar] [CrossRef] [Green Version]
  114. Zhuang, H.L.; Ganesh, P.; Cooper, V.R.; Xu, H.; Kent, P.R.C. Understanding the interactions between oxygen vacancies at SrTiO3 (001) surfaces. Phys. Rev. B 2014, 90, 064106. [Google Scholar] [CrossRef]
  115. Piskunov, S.; Heifets, E.; Eglitis, R.I.; Borstel, G. Bulk properties and electronic structure of SrTiO3, BaTiO3, PbTiO3 perovskites: An ab initio HF/DFT study. Comput. Mater. Sci. 2004, 29, 165–178. [Google Scholar] [CrossRef]
  116. Hay, P.J.; Wadt, W.R. Ab initio effective core potentials for molecular calculations. Potentials for K to Au including the outermost core orbitals. J. Chem. Phys. 1984, 82, 299–310. [Google Scholar] [CrossRef]
  117. Mayer, I. Bond Order and Valence: Relations to Mulliken’s Population Analysis. Int. J. Quantum Chem. 1984, 26, 151–154. [Google Scholar] [CrossRef]
  118. Bochicchio, R.C.; Reale, H.F. On the nature of crystalline bonding: Extension of statistical population analysis to two- and three- dimensional crystalline systems. J. Phys. B 1993, 26, 4871–4883. [Google Scholar] [CrossRef]
  119. Shi, H.; Chang, L.; Jia, R.; Eglitis, R.I. Ab initio calculations of the transfer and aggregation of F centers in CaF2. J. Phys. Chem. C 2012, 116, 4832–4839. [Google Scholar] [CrossRef]
  120. Mabud, S.A.; Glazer, A.M. Lattice parameters and birefringence in PbTiO3 single crystals. J. Appl. Cryst. 1979, 12, 49–53. [Google Scholar] [CrossRef]
  121. Edwards, J.W.; Speiser, R.; Johnston, H.L. Structure of Barium Titanate at Elevated Temperatures. J. Am. Chem. Soc. 1951, 73, 2934–2935. [Google Scholar] [CrossRef]
  122. Okazaki, A.; Scheel, H.J.; Müller, K.A. The lattice constant vs. temperature relation around the 105 K transition of a flux-grown SrTiO3 crystal. Phase Trans. 1985, 5, 207–218. [Google Scholar]
  123. Aoyagi, S.; Kuroiwa, Y.; Sawada, A.; Tanaka, H.; Nishibori, E.; Takata, M.; Sakata, M. Direct observation of covalency between O and disordered Pb in cubic PbZrO3. J. Phys. Soc. Jpn. 2002, 71, 2353–2356. [Google Scholar] [CrossRef]
  124. Benthem, K.; Elsässer, C.; French, R.H. Bulk electronic structure of SrTiO3: Experiment and theory. J. Appl. Phys. 2001, 90, 6156–6164. [Google Scholar] [CrossRef] [Green Version]
  125. Wemple, S.H. Polarization Fluctuations and the Optical-Absorption Edge in BaTiO3. Phys. Rev. B 1970, 2, 2679–2689. [Google Scholar] [CrossRef]
  126. Robertson, J. Band offsets of wide-band-gap oxides and implications for future electronic devices. J. Vacuum. Sci. Technol. B 2000, 18, 1785–1791. [Google Scholar] [CrossRef]
  127. Eglitis, R.I.; Popov, A.I. Systematic trends in (001) surface ab initio calculations of ABO3 perovskites. J. Saudi Chem. Soc. 2018, 22, 459–468. [Google Scholar] [CrossRef]
  128. Eglitis, R.I. Comparative First-Principles Calculations of SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001), (011) and (111) Surfaces. Ferroelectrics 2015, 483, 53–67. [Google Scholar] [CrossRef]
  129. Eglitis, R.I.; Vanderbilt, D. First-principles calculations of atomic and electronic structure of SrTiO3 (001) and (011) surfaces. Phys. Rev. B 2008, 77, 195408. [Google Scholar] [CrossRef]
  130. Eglitis, R.I.; Rohlfing, M. First-principles calculations of the atomic and electronic structure of SrZrO3 and PbZrO3 (001) and (011) surfaces. J. Phys.: Condens. Matter 2010, 22, 415901. [Google Scholar]
  131. Wang, Y.X.; Arai, M. First-principles study of the (001) surface of cubic SrZrO3. Surf. Sci. 2007, 601, 4092–4096. [Google Scholar] [CrossRef]
  132. Bickel, N.; Schmidt, G.; Heinz, K.; Müller, K. Ferroelectric relaxation of the SrTiO3 (100) surface. Phys. Rev. Lett. 1993, 62, 2009–2012. [Google Scholar] [CrossRef]
  133. Hikita, T.; Hanada, T.; Kudo, M.; Kawai, M. Structure and electronic state of the TiO2 and SrO terminated SrTiO3 (100) surfaces. Surf. Sci. 1993, 287, 377–381. [Google Scholar] [CrossRef]
  134. Piskunov, S.; Kotomin, E.A.; Heifets, E.; Maier, J.; Eglitis, R.I.; Borstel, G. Hybrid DFT calculations of the atomic and electronic structure for ABO3 perovskite (001) surfaces. Surf. Sci. 2005, 575, 75–88. [Google Scholar] [CrossRef]
  135. Eglitis, R.I. Ab initio calculations of CaZrO3, BaZrO3, PbTiO3 and SrTiO3 (001), (011) and (111) surfaces as well as their (001) interfaces. Integr. Ferroelectr. 2019, 196, 7–15. [Google Scholar] [CrossRef]
  136. Slassi, A.; Hammi, M.; Rhazouani, O.E. Surface Relaxations, Surface Energies and Electronic Structures of BaSnO3 (001) Surfaces: Ab initio Calculations. J. Electron. Mater. 2017, 46, 4133–4139. [Google Scholar] [CrossRef]
  137. Heifets, E.; Dorfman, S.; Fuks, D.; Kotomin, E. Atomistic simulation of the [001] surface structure in BaTiO3. Thin Solid Film. 1997, 296, 76–78. [Google Scholar] [CrossRef]
  138. Heifets, E.; Dorfman, S.; Fuks, D.; Kotomin, E.; Gordon, A. [001] Surface Structure in SrTiO3—Atomistic Study. Surf. Rev. Lett. 1998, 5, 341–345. [Google Scholar] [CrossRef]
  139. Sorokine, A.; Bocharov, D.; Piskunov, S.; Kashcheyevs, V. Electronic charge redistribution in LaAlO3 (001) thin films deposited at SrTiO3 substrate: First-principles analysis and the role of stoichiometry. Phys. Rev. B 2012, 86, 155410. [Google Scholar] [CrossRef]
Figure 1. Profile for the BO2-terminated (001) surface of ABO3 perovskite containing nine layers. The definitions of surface rumpling s as well as the near-surface interplanar separations Δd12 and Δd23 are explained.
Figure 1. Profile for the BO2-terminated (001) surface of ABO3 perovskite containing nine layers. The definitions of surface rumpling s as well as the near-surface interplanar separations Δd12 and Δd23 are explained.
Condensedmatter 07 00070 g001
Figure 2. Profile for the AO-terminated (001) surface of ABO3 perovskite containing nine layers.
Figure 2. Profile for the AO-terminated (001) surface of ABO3 perovskite containing nine layers.
Condensedmatter 07 00070 g002
Figure 3. Profile of the (001) interface among the cubic perovskites BaTiO3 and SrTiO3. SrTiO3 substrate planes are numbered using Arabic numbers. Planes of the augmented BaTiO3 (001) film are numbered using Roman numbers. Number 0 denotes the central (SrO) plane of the symmetrically TiO2-terminated 11-layer SrTiO3 (001) substrate.
Figure 3. Profile of the (001) interface among the cubic perovskites BaTiO3 and SrTiO3. SrTiO3 substrate planes are numbered using Arabic numbers. Planes of the augmented BaTiO3 (001) film are numbered using Roman numbers. Number 0 denotes the central (SrO) plane of the symmetrically TiO2-terminated 11-layer SrTiO3 (001) substrate.
Condensedmatter 07 00070 g003
Figure 4. Our B3PW computed electronic bulk band structure for BaTiO3 (a) and PbTiO3 (b) perovskites.
Figure 4. Our B3PW computed electronic bulk band structure for BaTiO3 (a) and PbTiO3 (b) perovskites.
Condensedmatter 07 00070 g004
Figure 5. Our B3PW computed band gaps for TiO2-terminated (001) surfaces for: (a) BaTiO3 perovskite and (b) PbTiO3 perovskite.
Figure 5. Our B3PW computed band gaps for TiO2-terminated (001) surfaces for: (a) BaTiO3 perovskite and (b) PbTiO3 perovskite.
Condensedmatter 07 00070 g005
Figure 6. Our B3PW computed band gaps for AO-terminated (001) surfaces for: (a) BaTiO3 perovskite and (b) PbTiO3 perovskite.
Figure 6. Our B3PW computed band gaps for AO-terminated (001) surfaces for: (a) BaTiO3 perovskite and (b) PbTiO3 perovskite.
Condensedmatter 07 00070 g006
Figure 7. Our hybrid-DFT computed upper-layer atomic displacement magnitudes Δz (% of joint lattice constant a0) for an 11-layer STO (001) substrate (x = 0) and for 10 augmented BTO layers (x = 1–10).
Figure 7. Our hybrid-DFT computed upper-layer atomic displacement magnitudes Δz (% of joint lattice constant a0) for an 11-layer STO (001) substrate (x = 0) and for 10 augmented BTO layers (x = 1–10).
Condensedmatter 07 00070 g007
Figure 8. Our hybrid-DFT computed upper-layer atomic displacement magnitudes Δz (% of joint lattice constant a0) for the 11-layer STO (001) substrate (x = 0) and for 10 augmented PTO layers (x= 1–10).
Figure 8. Our hybrid-DFT computed upper-layer atomic displacement magnitudes Δz (% of joint lattice constant a0) for the 11-layer STO (001) substrate (x = 0) and for 10 augmented PTO layers (x= 1–10).
Condensedmatter 07 00070 g008
Figure 9. Our hybrid-DFT computed Γ-Γ band gap (in eV) for the PTO/STO (001) interface depending on the number of PTO (001) layers (x = 1–10) augmented on the 11-layer symmetrical TiO2-terminated SrTiO3 (001) substrate.
Figure 9. Our hybrid-DFT computed Γ-Γ band gap (in eV) for the PTO/STO (001) interface depending on the number of PTO (001) layers (x = 1–10) augmented on the 11-layer symmetrical TiO2-terminated SrTiO3 (001) substrate.
Condensedmatter 07 00070 g009
Figure 10. Our B3PW computed Γ-Γ band gap (in eV) for BTO/STO (a) and PZO/SZO (b) (001) interfaces, containing an 11-layer STO or SZO substrate as well as from 1 to 10 augmented BTO or PZO layers, respectively. Dashed lines serve as a guide for the eyes.
Figure 10. Our B3PW computed Γ-Γ band gap (in eV) for BTO/STO (a) and PZO/SZO (b) (001) interfaces, containing an 11-layer STO or SZO substrate as well as from 1 to 10 augmented BTO or PZO layers, respectively. Dashed lines serve as a guide for the eyes.
Condensedmatter 07 00070 g010
Table 1. Our hybrid-DFT computed effective atomic charges Q (in e) as well as bond populations P (in e) for atoms in PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 bulk matrices.
Table 1. Our hybrid-DFT computed effective atomic charges Q (in e) as well as bond populations P (in e) for atoms in PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 bulk matrices.
Bulk PropertiesPbTiO3BaTiO3SrTiO3PbZrO3SrZrO3
IonPropertyB3PWB3PWB3PWB3LYPB3LYP
AQ+1.354+1.797+1.871+1.368+1.880
P+0.016−0.034−0.010+0.030+0.002
OQ−1.232−1.388−1.407−1.160−1.351
P+0.098+0.098+0.088+0.106+0.092
BQ+2.341+2.367+2.351+2.111+2.174
Table 2. Our B3PW computation results as well as the experimental data for SrTiO3, BaTiO3 and PbTiO3 bulk Γ-Γ band gaps (in eV) and their structure at room temperature (RT).
Table 2. Our B3PW computation results as well as the experimental data for SrTiO3, BaTiO3 and PbTiO3 bulk Γ-Γ band gaps (in eV) and their structure at room temperature (RT).
SolidStructure at RTExp. Γ-Γ Band Gap (eV)Ab initio Γ-Γ Band Gap
SrTiO3Cubic phase3.75 eV [124]3.96 eV
BaTiO3Tetragonal ↔ orthorhombic phase at 278 K 3.38 eV (//c);
3.27 eV (⊥ c) [125]
3.55 eV
PbTiO3Tetragonal phase3.4 eV [126]4.32 eV
Table 3. Atomic relaxation magnitudes (in % of the a0) for the BO2-terminated PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 (001) surfaces.
Table 3. Atomic relaxation magnitudes (in % of the a0) for the BO2-terminated PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 (001) surfaces.
SolidPbTiO3BaTiO3SrTiO3PbZrO3SrZrO3
LayerIonTiO2-t.TiO2-t.TiO2-t.ZrO2-t.ZrO2-t.
1B−2.81−3.08−2.25−2.37−1.38
O+0.31−0.35−0.13−1.99−2.10
2A+5.32+2.51+3.55+4.36+2.81
O+1.28+0.38+0.57+1.04+0.91
3B---−0.47−0.04
O---−0.28−0.05
Table 4. Atomic relaxation magnitudes (in % of the a0) for the AO-terminated PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 (001) surfaces.
Table 4. Atomic relaxation magnitudes (in % of the a0) for the AO-terminated PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 (001) surfaces.
SolidPbTiO3BaTiO3SrTiO3PbZrO3SrZrO3
LayerIonPbO-t.BaO-t.SrO-t.PbO-t.SrO-t.
1A−3.82−1.99−4.84−5.69−7.63
O−0.31−0.63+0.84−2.37−0.86
2B+3.07+1.74+1.75+0.57+0.86
O+2.30+1.40+0.77+0.09−0.05
3A---−0.47−1.53
O---−0.47−0.45
Table 5. Our hybrid-DFT computed surface rumpling s as well as relative displacements Δdij for BO2-terminated PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 (001) surfaces in (% of a0). Relevant experimental data are listed in Table 5 for comparison purposes.
Table 5. Our hybrid-DFT computed surface rumpling s as well as relative displacements Δdij for BO2-terminated PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrZrO3 (001) surfaces in (% of a0). Relevant experimental data are listed in Table 5 for comparison purposes.
SolidFunctionalTerminationsΔd12Δd23
PbZrO3B3LYPZrO2+0.38−6.73+4.83
SrZrO3B3LYPZrO2−0.72−4.19+2.85
LDA [131]ZrO2−0.7−6.1+4.2
PbTiO3B3PWTiO2-term.+3.12−8.13+5.32
BaTiO3B3PWTiO2-term.+2.73−5.59+2.51
SrTiO3B3PWTiO2-term.+2.12−5.79+3.55
LEED [132]TiO2-term.+2.1 ± 2+1 ± 1−1 ± 1
RHEED [133]TiO2-term.+2.6+1.8+1.3
Table 6. Our hybrid-DFT computed atomic displacement absolute magnitudes D (in Å) and the effective charges Q (in e), as well as the chemical bond populations P (in e), for BO2-terminated PbZrO3, SrZrO3, PbTiO3, BaTiO3 and SrTiO3 (001) surfaces.
Table 6. Our hybrid-DFT computed atomic displacement absolute magnitudes D (in Å) and the effective charges Q (in e), as well as the chemical bond populations P (in e), for BO2-terminated PbZrO3, SrZrO3, PbTiO3, BaTiO3 and SrTiO3 (001) surfaces.
PbZrO3SrZrO3PbTiO3BaTiO3SrTiO3
LayerProp.IonZrO2-t.ZrO2-t.TiO2-t.TiO2-t.TiO2-t.
1DB−0.100−0.058−0.111−0.123−0.088
Q+2.165+2.196+2.279+2.307+2.291
P+0.116+0.114+0.114+0.126+0.118
DO−0.084−0.088+0.012−0.014−0.005
Q−1.171−1.277−1.184−1.280−1.296
P+0.046−0.002+0.044−0.038−0.014
2DA+0.184+0.118+0.209+0.101+0.139
Q+1.357+1.869+1.275+1.767+1.850
P+0.022+0.002+0.008−0.030−0.008
DO+0.044+0.038+0.050+0.015+0.022
Q−1.103−1.287−1.167−1.343−1.365
P+0.098+0.094+0.080+0.090+0.080
3DB−0.020−0.001---
Q+2.116+2.172+2.335+2.365+2.348
P+0.124+0.102+0.108+0.104+0.096
DO−0.012−0.002---
Q−1.148−1.331−1.207−1.371−1.384
P+0.036+0.002+0.018−0.034−0.010
Table 7. Our hybrid-DFT computed cleavage and relaxation, as well as surface energies, for PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrTiO3 (001) surfaces (in eV per surface cell).
Table 7. Our hybrid-DFT computed cleavage and relaxation, as well as surface energies, for PbTiO3, BaTiO3, SrTiO3, PbZrO3 and SrTiO3 (001) surfaces (in eV per surface cell).
SurfaceTerminationEcleavageErelaxationEsurface
PbTiO3 (001)TiO21.02−0.280.74
PbO−0.190.83
BaTiO3 (001)TiO21.30−0.231.07
BaO−0.111.19
SrTiO3 (001)TiO21.39−0.161.23
SrO−0.241.15
PbZrO3 (001)ZrO21.20−0.270.93
PbO−0.201.00
SrZrO3 (001)ZrO21.56−0.321.24
SrO−0.431.13
Table 8. Our hybrid-DFT computed direct Γ-Γ band gaps (in eV) for PbTiO3, BaTiO3 and SrTiO3 perovskite bulk, as well as their BO2 and AO-terminated (001) surfaces. The experimental PbTiO3, BaTiO3 and SrTiO3 bulk Γ-Γ band gaps are listed for comparison purposes.
Table 8. Our hybrid-DFT computed direct Γ-Γ band gaps (in eV) for PbTiO3, BaTiO3 and SrTiO3 perovskite bulk, as well as their BO2 and AO-terminated (001) surfaces. The experimental PbTiO3, BaTiO3 and SrTiO3 bulk Γ-Γ band gaps are listed for comparison purposes.
TerminationPbTiO3 (Γ-Γ) GapBaTiO3 (Γ-Γ) GapSrTiO3 (Γ-Γ) Gap
Bulk 4.32 eV3.55 eV3.96 eV
TiO2-term. (001)3.18 eV2.96 eV3.95 eV
AO-term. (001)3.58 eV3.49 eV3.72 eV
Experiment3.4 eV [134] 3.2 eV [125]3.75 eV [124]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Eglitis, R.I.; Piskunov, S.; Popov, A.I.; Purans, J.; Bocharov, D.; Jia, R. Systematic Trends in Hybrid-DFT Computations of BaTiO3/SrTiO3, PbTiO3/SrTiO3 and PbZrO3/SrZrO3 (001) Hetero Structures. Condens. Matter 2022, 7, 70. https://doi.org/10.3390/condmat7040070

AMA Style

Eglitis RI, Piskunov S, Popov AI, Purans J, Bocharov D, Jia R. Systematic Trends in Hybrid-DFT Computations of BaTiO3/SrTiO3, PbTiO3/SrTiO3 and PbZrO3/SrZrO3 (001) Hetero Structures. Condensed Matter. 2022; 7(4):70. https://doi.org/10.3390/condmat7040070

Chicago/Turabian Style

Eglitis, Roberts I., Sergei Piskunov, Anatoli I. Popov, Juris Purans, Dmitry Bocharov, and Ran Jia. 2022. "Systematic Trends in Hybrid-DFT Computations of BaTiO3/SrTiO3, PbTiO3/SrTiO3 and PbZrO3/SrZrO3 (001) Hetero Structures" Condensed Matter 7, no. 4: 70. https://doi.org/10.3390/condmat7040070

Article Metrics

Back to TopTop