ACS Publications. Most Trusted. Most Cited. Most Read
Combined Theoretical and Experimental Study of the Moiré Dislocation Network at the SrTiO3-(La,Sr)(Al,Ta)O3 Interface
My Activity

Figure 1Loading Img
  • Open Access
Functional Inorganic Materials and Devices

Combined Theoretical and Experimental Study of the Moiré Dislocation Network at the SrTiO3-(La,Sr)(Al,Ta)O3 Interface
Click to copy article linkArticle link copied!

  • Chiara Ricca
    Chiara Ricca
    Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012 Bern, Switzerland
    More by Chiara Ricca
  • Elizabeth Skoropata
    Elizabeth Skoropata
    Swiss Light Source, Paul Scherrer Institut, Forschungsstrasse 111, 5232 Villigen PSI, Switzerland
  • Marta D. Rossell
    Marta D. Rossell
    Electron Microscopy Center, Empa, Swiss Federal Laboratories for Materials Science and Technology, Überlandstrasse 129, 8600 Dübendorf, Switzerland
  • Rolf Erni
    Rolf Erni
    Electron Microscopy Center, Empa, Swiss Federal Laboratories for Materials Science and Technology, Überlandstrasse 129, 8600 Dübendorf, Switzerland
    More by Rolf Erni
  • Urs Staub
    Urs Staub
    Swiss Light Source, Paul Scherrer Institut, Forschungsstrasse 111, 5232 Villigen PSI, Switzerland
    More by Urs Staub
  • Ulrich Aschauer*
    Ulrich Aschauer
    Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012 Bern, Switzerland
    Department of Chemistry and Physics of Materials, University of Salzburg, Jakob-Haringer-Street 2A, A-5020 Salzburg, Austria
    *Email: [email protected]
Open PDF

ACS Applied Materials & Interfaces

Cite this: ACS Appl. Mater. Interfaces 2023, 15, 46, 53678–53687
Click to copy citationCitation copied!
https://doi.org/10.1021/acsami.3c10958
Published November 9, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

Recently, a highly ordered Moiré dislocation lattice was identified at the interface between a SrTiO3 (STO) thin film and the (LaAlO3)0.3(Sr2TaAlO6)0.7 (LSAT) substrate. A fundamental understanding of the local ionic and electronic structures around the dislocation cores is crucial to further engineer the properties of these complex multifunctional heterostructures. Here, we combine experimental characterization via analytical scanning transmission electron microscopy with results of molecular dynamics and density functional theory calculations to gain insights into the structure and defect chemistry of these dislocation arrays. Our results show that these dislocations lead to undercoordinated Ta/Al cations at the dislocation core, where oxygen vacancies can easily be formed, further facilitated by the presence of cation vacancies. The reduced Ti3+ observed experimentally at the dislocations by electron energy-loss spectroscopy is a consequence of both the structure of the dislocation itself and of the electron doping due to oxygen vacancy formation. Finally, the experimentally observed Ti diffusion into the LSAT around the dislocation core occurs only together with cation vacancy formation in the LSAT or Ta diffusion into STO.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2023 The Authors. Published by American Chemical Society

1. Introduction

Click to copy section linkSection link copied!

Complex transition metal perovskite oxides are a versatile class of materials with a wide spectrum of functional properties. They can be insulating, semiconducting, or metallic and show technologically relevant phenomena such as magnetism, ferroelectricity, or the more exotic high-temperature superconductivity and colossal magnetoresistance. (1−4) These properties are the result of a complex interplay of charge, orbital, spin, and lattice degrees of freedom (1) and depend strongly on strain and the defect chemistry. (5−7) The structural compatibility between different perovskites allows them to be stacked on top of each other, and the advances in deposition techniques enabled the fabrication of complex multifunctional heterostructures with relative ease. These heterostructures often give rise to interesting, novel, and unexpected physical phenomena emerging at the interface where materials with different structural and electronic properties meet: quasi-two-dimensional (2D) electron gas, colossal ionic conductivity, giant thermoelectric effect, or resistance switching. (1,3,8−12)
Recently, a highly ordered Moiré lattice has been identified at the interface between SrTiO3 (STO) and (LaAlO3)0.3(Sr2TaAlO6)0.7 (LSAT) by high-resolution X-ray diffraction reciprocal space mapping. (13) A 30 nm thick film of STO was grown on LSAT (001) by pulsed laser deposition, followed by 12 h of annealing at 1200 °C and ambient pressure. STO (aSTO = 3.905 Å) and LSAT (aLSAT = 3.869 Å) both have a cubic lattice with a small mismatch of 0.93%. The high-temperature annealing allows the almost complete relaxation of the STO film, which results in the appearance of the Moiré pattern. This newly formed 2D pattern has a Moiré lattice constant of approximately 40 nm, corresponding to 106/107 unit cells of STO/LSAT, necessary to compensate for the small lattice mismatch between the two materials. Scanning transmission electron microscopy (STEM) images suggest that this periodicity is related to the appearance of a network of edge dislocations at the interface that has the same periodicity as the Moiré pattern.
The ability to form such ordered superlattices with a 2D network of line defects at complex perovskite oxide interfaces could be an emerging avenue to induce new interfacial functionalities with unforeseen potential applications, such as 2D ferroics, 2D grid conductivity along the defect lines, or ferroelectric three-dimensional (3D) vortex structures. These interfacial phenomena are a consequence of spin and charge interactions at the interface, which are in turn controlled by the local atomic arrangement. Hence, a detailed understanding of the atomic structure and electronic properties of the STO/LSAT interface in the presence of dislocations is key to interpreting the behavior of this heterostructure and to enhance its functional properties.
Given that the oxide defect chemistry and diffusion are known to be heavily affected by strain (14−17) and since dislocations are surrounded by strong strain fields, one would expect preferential defect formation and segregation around dislocations. Indeed, recent computational studies have focused not only on describing the structure of interfacial misfit dislocations (18,19) but also on their interaction with dopants, defects, and defect clusters. (20−22) These classical potential-based calculations, however, always treat charged point defects since whole ions are removed. Electronic structure density functional theory (DFT) calculations would be necessary to account for the formation of neutral or not fully ionic defects. In the present case, DFT is, however, limited by the large size of the supercells required to describe the periodicity of the Moiré lattice and misfit dislocations. In addition, the huge configurational space of point defects and point-defect clusters in the symmetry-broken environment represents a challenge, given the computational cost of DFT.
To overcome these challenges, we use classical molecular dynamics (MD) to obtain a dislocation structure with the correct long-range strain field of the Moiré interface with misfit dislocations and combine it with DFT calculations on cluster models extracted from these classically relaxed structures. The space of possible defects and their location around the dislocation are reduced by experimental information obtained from chemical mapping by electron energy-loss (EELS) and energy-dispersive X-ray (EDX) spectroscopies. The results lead to a detailed understanding of the atomic-scale defect structure and the resulting electronic properties around the interfacial dislocation network that will guide the design and optimization of these promising heterostructures.

2. Methods

Click to copy section linkSection link copied!

2.1. Computational Details

2.1.1. Dislocation Models

The simulation of these dislocation arrays is complicated by their nonperiodic nature along the interface normal and by their large ordering period along the interface. An appropriate description along both of these directions is, however, crucial to accurately determine the long-range strain field associated with the dislocation. The interfacial Moiré periodicity can be determined by X-ray diffraction experiments (see Figure 1a–c for a plane-view and line cuts through the reciprocal space volume) and best fits simulated diffraction patterns for a 106/107 type Moiré motif. (13) This periodicity with a Moiré lattice constant of 40.8 nm also compensates for the 0.93% lattice mismatch between the STO film and the LSAT substrate. The associated interfacial dislocation structure matches this periodicity as shown by the high-angle annular dark-field (HAADF)-STEM image and associated strain map in Figure 1d,e.

Figure 1

Figure 1. (a) In-plane projection of the X-ray diffraction of the (22̅4) reflection of SrTiO3/LSAT thin films with a Moiré lattice. Line cuts through the (b) film (L = 4.000) and (c) substrate (L = 4.034) reflections show the periodicity of the Moiré lattice. (d) HAADF-STEM image of a STO thin film grown on a LSAT (001) substrate and (e) corresponding strain map along the x direction calculated by geometric phase analysis. The misfit dislocations appear as butterflylike features at the STO/LSAT interface with a compression region (in blue) and a tensile region (in yellow). (f) Structure of the periodically repeated relaxed MD model of the dislocation array at the SrTiO3/LSAT interface with the dislocations and the (2 × 107 × 60) simulation cell highlighted. Only the B–O bonding network and the A-site atoms are shown for clarity. Insets show the initial unrelaxed part of the structure with the removed STO plane and the extracted DFT cluster model, respectively. In the DFT model, the faded-out atoms are fixed during DFT geometry optimization to impose elastic boundary conditions.

In order to obtain a reliable computational description of the 2D Moiré pattern at the interface of a thin film of STO on LSAT (001), we employed the simulation setup schematically shown in Figure 1f, which is a combination of the simulation approaches suggested in refs (20), (23), and (24). We started by creating two dislocations with opposite Burgers vectors in a simulation box with dimensions ∼0.75 nm × 41.50 nm × 23.51 nm (approximately 64,000 atoms) corresponding to a 2 × 107 × 60 supercell of a 5-atom cubic perovskite unit cell. The number of cells along the y axis allows us to reproduce the observed periodicity of the Moiré superlattice, while along the z axis, we stacked 30 STO on 30 LSAT layers. This means that the two dislocation cores are separated by about 11.76 nm along the z axis, which is sufficient to significantly suppress interactions between them. A missing plane is then created in the middle of the cell by removing one SrO and one TiO2 plane (see Figure 1f). The system is relaxed via classical MD (see below for details), during which the missing plane heals, forming two dislocations with opposite Burgers vectors at the upper and lower interfaces.
DFT calculations (see below for details) are then performed on a cluster model created from the final MD structure by extracting a section of 2 × 10 × 10 STO/LSAT unit cells around the bottom dislocation core (Figure 1f). This cell contains 950 atoms and 5 STO and 5 LSAT layers along the z axis. For the LSAT substrate, 18 La and 82 Sr, as well as 59 Al and 41 Ta atoms, were randomly introduced at the A and B sites, respectively, to obtain a charge-neutral LSAT layer with a composition (La0.2Sr0.8Al0.4Ta0.6O6) similar to that of the experiment (La0.3Sr0.7Al0.3Ta0.7O6). The cluster was periodically repeated along the dislocation line only (x axis), while in the perpendicular directions, only the periodicity of the electrostatic potential was imposed. Furthermore, atoms in the boundary region were kept fixed throughout the simulation to impose the elastic boundary conditions of the interface and dislocation environment. Within this approach, it is possible to avoid artifacts due to the interaction of the dislocation with its images in neighboring cells, thus isolating the dislocation core. The dislocation core is embedded in the correct long-range elastic field, allowing, at the same time, to minimize any spurious effect due to long-range electrostatic fields. (23)
Neutral vacancies (VX) were created by removing one X (X = O, La, Al, and Ta) atom from the DFT dislocation model. Ti-substitution at Al (TiAl) or Ta (TiTa) sites and Ta-substitution of Ti atoms (TaTi) were also taken into account. Different possible vacancies and substitutional defect configurations involving different sites around the dislocation were considered. All data are available on the Materials Cloud Archive. (25)

2.1.2. Molecular Dynamics

The LAMMPS (26) code was used to perform the classical MD simulations. Interatomic interactions were described using a nonpolarizable rigid-ion model consisting of long-range electrostatic interactions between the nuclei and a short-range Buckingham potential with a cutoff radius of 12 Å
Vij=qiqj4πϵ0rij+Aijerij/ρijCijrij6
(1)
where qi/j is the atomic charge, rij is their separation, and Aij, ρij, and Cij are the parameters of the short-range potential as reported in Table 1. The short-range parameters were derived by starting from the Lewis and Catlow (27) set and using the GULP code (28−30) to fit lattice parameters and elastic constants of STO and LSAT derived from DFT calculations (see below). For LSAT, we chose to describe the interaction between the O and A/B dummy atoms representing the average properties of the La/Sr and Al/Ta atoms occupying the A and B sites, respectively. MD simulations were performed in the canonical (NVT) ensemble with a Nose–Hoover thermostat and barostat. The system was allowed to relax first at 50 K for 50 ps and then at 1473 K for 130 ps before the temperature was reduced again to 50 K over 120 ps. In order to drain the kinetic energy released during closing the missing plane in a controlled way, a viscous damping force with a coefficient γ = 1.0 eV·ps/Å2 is applied to all atoms.
Table 1. Parameters of the Coulomb–Buckingham Potentials for STO and LSAT
ijqi (e)Aij (eV)ρij (Å)Cij (eV/Å6)
SrO2.001324.770.30080.00
TiO4.00762.260.40140.00
AO2.182018.140.28760.00
BO3.82867.030.38280.00
OO–2.0022764.300.149031.15
DFT calculations to determine lattice parameters and elastic constants used in potential fitting were performed with the VASP code. (31−34) We used the PBE (35) exchange-correction functional together with PAW (36,37) potentials with La(5s, 5d, 5p, and 6s), Sr(4s, 4p, and 5s), Al(3s and 3p), Ta(5p, 5d, and 6s), and O(2s and 2p) valence electrons and a plane-wave cutoff of 550 eV. For LSAT, we used a 7.76Å×10.97Å×10.97Å2×22×22 supercell of the 5-atom cubic unit cell with composition La3Sr13Al9Ta7O48, the reciprocal space of which was sampled with a 4 × 4 × 4 Monkhorst–Pack (38) mesh, while a 8 × 8 × 8 mesh was used for the 5-atom STO unit cell (3.90 Å × 3.90 Å × 3.90 Å). Structures were relaxed until forces converged below 10–4 eV/Å before elastic constants were determined using a central finite-differences approach with a step size of 0.015 Å.

2.1.3. DFT Calculations

The DFT calculations on the cluster model were performed with the CP2K program package (39) using the PBE (35) exchange–correlation functional. The norm-conserving Goedecker–Teter–Hutter (GTH) pseudopotentials (40) together with the GTH double-ζ polarized molecularly optimized basis sets (41) and an energy cutoff of 750 Ry were applied. The convergence criterion for the self-consistent field method was set to 1 × 10–6 Ha, while atomic positions were relaxed within a force threshold of 10–3 eV/Å.
The defect formation energy (Ef) of a neutral defect was calculated as described in ref (42)
Ef(μi)=Etot,defEtot,stoiciniμi
(2)
where Etot,def and Etot,stoic are the DFT total energies of the defective system and of the stoichiometric cell, respectively. ni indicates the number of atoms of a certain species i that is added (ni > 0) or removed (ni < 0) from the supercell to form the defect, while μi is the species’ chemical potential. For simplicity, we used the atomic energy of the corresponding reference phase for each element (metal La, Ta, Ti, Al, and molecular O2) as the chemical potential.

2.2. Experimental Methods

2.2.1. X-ray Diffraction

X-ray diffraction was measured at the surface diffraction endstation of the materials science beamline at the Swiss Light Source. X-rays of 12.65 keV were focused with a beam size of 500 × 500 μm on the sample at room temperature. The (22̅4) reflection was measured with an area detector after alignment of the UB matrix with 3 orthogonal and 2 nonorthogonal reflections. The detector images were converted to reciprocal space according to Schlepütz et al., (43) and the data were analyzed using the xrayutilities package. (44)

2.2.2. Scanning Transmission Electron Microscopy

Electron-transparent samples for STEM investigations were produced in a cross-section geometry using an FEI Helios 660 G3 UC dual-beam focused (Ga) ion beam instrument operated at 30 and 5 kV, after the deposition of C and Pt protective layers. HAADF STEM, EELS, and EDX were carried out using a probe aberration-corrected FEI Titan Themis microscope operated at 300 kV and equipped with a SuperX EDX system and a CEFID energy filter in combination with an ELA direct electron detector (for details, see ref (45)). For the HAADF-STEM data acquisition, a probe convergence semiangle of 26 mrad was set, and the inner angle of the annular semidetection range was 171 mrad. The EELS data were obtained with a collection semiangle of 35 mrad yielding an effective collection angle of about 29.5 mrad considering the energy range of the spectra.
A quantitative analysis of the lattice distortions at the dislocation cores was performed by means of peak-pair analysis (PPA), fitting the peaks corresponding to the atomic columns in the high-resolution HAADF-STEM images as described in ref (46). In particular, we analyzed the structural distortions by measuring the distance between the peaks corresponding to the atomic columns of the A sublattice.

3. Results and Discussion

Click to copy section linkSection link copied!

3.1. Structure and Strain Field at the Dislocation Core

As already discussed in ref (13), high-resolution STEM images of the annealed STO/LSAT samples reveal the presence of a highly ordered arrangement of edge dislocations with periodicity equivalent to the one of the Moiré lattice observed by high-resolution X-ray diffraction reciprocal space mapping. These dislocations form where the local structural mismatch between the two materials is the largest (see Figure 2b). In the present work, the HAADF-STEM images have been used to map the strain field around the dislocation core, as reported in Figure 2a. The strain field is, indeed, highly localized around the dislocation core, where two different strain regions can be identified: a tensile strain region in the STO film and a compressive strain region in the LSAT layer.

Figure 2

Figure 2. (a) Map of the interatomic distances (in Å) for the A sublattice along the horizontal direction as obtained by PPA from the HAADF-STEM image in (b). (c) Map of horizontal interatomic A–A distances extracted from the DFT-relaxed dislocation core structure shown in (d). Circles are located at the midpoint of A–A pairs and are color-coded according to the distance between that pair. (e) Electronic density of states (DOS) projected on the atoms at the dislocation core for the stoichiometric dislocation model. The vertical dashed line indicates the position of the Fermi level. (f) Charge density isosurface (10–2 e/Å3) in the energy range highlighted in red in (e).

Figure 2d shows the final structure of the dislocation core obtained after DFT geometry optimization starting from the structure extracted from the MD simulations. The strain field associated with this model was computed by extracting A–A distances parallel to the interface, as shown in Figure 2c. In this case, the largest strains are also observed at the dislocation core with an average tensile/compressive strain in the STO/LSAT layer, respectively. Overall, the strain-field maps derived from experiments and DFT calculations show a qualitative agreement, allowing us to confidently use the DFT model to understand the structural properties of the dislocation core at the atomic level.
As shown in Figure 2d, highly strained Ti–O–Ti bonds are established across the missing planes, leaving the B cations of the extra Ta/AlO2 plane undercoordinated at the dislocation core. The missing Al–O bonds at the STO/LSAT interface induce hole doping of the system, as can be seen from the DOS projected on the atoms at the dislocation core (Figure 2e), where the Fermi level crosses the top of the valence band, which is formed by the O-2p states of the O atoms bonded to the undercoordinated B cations at the dislocation core (Figure 2f).
Despite the overall qualitative agreement, the dislocation core imaged by STEM shows contrast differences larger than expected from the structural and electronic changes in the DFT model, which could result from the DFT model being constructed with stoichiometric STO and LSAT, while point defects, such as oxygen and cation vacancies or substitutional defects, may be present in experiment.

3.2. Defect Chemistry of the Dislocation Core

3.2.1. Oxygen Vacancies

EELS applied to the STO/LSAT interface can be used to obtain information on both the O and Ti states from the O–K (O 1s → 2p) and Ti–L3,2 (Ti 2p → 3d) core edges and could thus provide information about the presence of oxygen vacancies at the interface between the two oxides. The O–K edge map of Figure 3b appears to be darker at the dislocation core, pointing to the presence of oxygen vacancies in this region. At the O–K edge, the unoccupied O p-DOS in the presence of a core hole is probed, and thus, its intensity is expected to be reduced in the presence of oxygen vacancies that are, generally, electron donors in transition metal oxides. (47)

Figure 3

Figure 3. (a) HAADF-STEM image of a dislocation at the STO/LSAT interface and corresponding atomic-resolution maps of the (b) O–K and (c) Ti–L3,2 excitation edges. The scale bar is 1 nm. (d) Representative Ti–L3,2 EELS spectra of the STO film and the surroundings of the dislocation core. The spectra have been normalized to the L2 eg peak height for clarity.

To investigate the formation of oxygen vacancies (VO) around the dislocation using DFT, we considered, for simplicity, a single neutral VO created at different oxygen sites in our STO/LSAT dislocation model. The map reported in Figure 4a is color-coded according to the VO formation energy of each VO site relative to that of the most stable site (ΔEf). The energetically most favorable site to form a VO is at the dislocation core for the O atoms of the extra Ta/AlO2 plane closest to the STO/LSAT interface. This can be ascribed to the removal of an undercoordinated O atom to form this defect. It was already observed that similar to what happens at surfaces and grain boundaries, also at the dislocation core, undercoordination can result in a lowering of Ef. (20) Furthermore, as seen above, the DOS of the stoichiometric dislocation model is characterized by hole states right above the Fermi level and localized on the undercoordinated O atoms at the dislocation core (Figure 2e). These hole states can be filled by the two extra electrons left in the lattice upon the formation of a neutral VO (Figure 4b). The formation energy gradually increases with increasing distance from the dislocation core, with larger values in LSAT compared to those in STO, which is easier to reduce, as shown by the computed VO formation energies in the two bulk materials (STO: 3.90 eV and LSAT: 5.04 eV). It also agrees with tensile strain, as in the STO layer, favoring neutral VO formation based on chemical expansion arguments. (48) These results suggest that VO has the tendency to segregate to the dislocation core and, in particular, to the STO side of the interface, in agreement with information obtained from the O–K edge EELS spectral map.

Figure 4

Figure 4. (a) Color map of the relative VO formation energy (ΔEf) around the dislocation core. Electronic DOS projected on the atoms at the dislocation core for the STO/LSAT interface model containing (b) one, (c) two, and (d) three VO in the most stable configurations. The vertical dashed line indicates the position of the Fermi level.

The Ti-L3,2 map of Figure 3c shows that Ti atoms tend to diffuse into the LSAT. Interestingly, the Ti-L3,2 EELS spectra acquired in the STO film far away from the interface and around the dislocation core show striking differences (Figure 3d). Both spectra are composed of two main features, namely, the L3 and L2 edges, separated by about ∼5.5 eV due to the spin–orbit splitting of the Ti 2p core hole into 2p3/2 and 2p1/2 states. Besides, these edges are further subdivided into two peaks, the t2g and eg peaks, by the strong octahedral crystal-field splitting arising from the surrounding oxygen atoms. In particular, the spectrum acquired at the dislocation is characterized by a relative increase in the spectral weight of the Ti-L3,2 t2g peaks, especially in the higher-energy L2 edge, compared with the spectrum of the STO film. Previously, such spectral changes were related to the presence of Ti3+ in nominally Ti4+-based perovskite oxides. (49) Aside from this, the energy shift to lower energies of the Ti-L3,2 edge observed in the dislocation spectrum with respect to the spectra obtained in the STO film suggests the presence of reduced Ti3+ species in the dislocation core compared to Ti4+ in the STO film. (47)
This change in the oxidation state is consistent with electron doping due to neutral VO at the dislocation core. Unfortunately, our DFT results did not show a complete reduction of Ti atoms at or around the dislocation core. Indeed, the DOS in Figure 4b–d do not show any filled localized state with Ti-3d character, inherent with reduction of Ti atoms to Ti3+ even in the presence of three VO, positioned, for simplicity, at the most stable sites according to Figure 4a. The absence of DOS features related to Ti reduction could be due to the limitations of our DFT method. It is well known that standard semilocal DFT functionals, such as PBE, fail in localizing the charge on Ti atoms neighboring a VO for small bulk STO cells. (50) Test calculations using a Hubbard-corrected PBE + U functional did not yield localized Ti3+ either. We note that hybrid functionals have shown promise in obtaining these excess charge locations in STO. (51−53) Apart from being too computationally intensive for our simulation cells, these functionals may, due to the slight overestimation of the band gap, result in states that are too deep in the gap and, given the underestimation of the crystal field splitting, may have an eg rather than a t2g character. (54,55) The position of the defect states is, however, improved for very large cell sizes, where oxygen vacancy defect states become shallower. (54)
However, with increasing the number of VO, we observe a small peak with Ti-3d character at the bottom of the CB (see the arrow in Figure 4b) that is present for a single VO to disappear, while the states with O-2p character at the top of the valence band are lowered in energy. This suggests a filling of Ti states for an increasing number of VO and hence an increase in electron doping. This is supported by an analysis of the Mulliken charges for Ti atoms. Already in the stoichiometric dislocation model, Ti ions close to the dislocation core are more reduced compared to the rest of the Ti sites in STO (Figure 5a). The creation of the first VO at the dislocation core, when the dislocation hole state is only partially filled, does not significantly alter the charge of these Ti atoms. These charges, however, increase for 2 or 3 VO in STO around the dislocation (Figures 5b–d). These results suggest that the Ti3+ detected by EELS in the vicinity of the dislocation stems from both the presence of the dislocation and the formation of VO.

Figure 5

Figure 5. Mulliken charges for the Ti atoms in (a) the stoichiometric dislocation model and in the presence of (b) one, (c) two, and (d) three VO. The black squares indicate the position of the VO.

3.2.2. Cation Vacancies

To further characterize the defect chemistry of the STO/LSAT interface, we performed chemical mapping with atomic resolution using EDX. The positions of the atomic columns visualized in the EDX maps, shown in Figure 6b–g for Sr, Ti, O, La, Al, and Ta, reveal the chemical structure at the STO/LSAT interface at and around the dislocation core: cation vacancies (especially VAl, VLa, and VTa) are formed around the dislocation core, which are (partially) filled by Ti diffusing from STO into LSAT. At the same time, a partial substitution of Ti by Ta takes place in the STO film above the dislocation.

Figure 6

Figure 6. (a) HAADF-STEM image of a misfit dislocation at the STO/LSAT interface and (b–g) corresponding elemental maps of Sr, Ti, O, La, Al, and Ta calculated from an EDX spectrum image using the Sr–K, Ti–K, O–K1, La–L, Al–K1, and Ta–L lines, respectively. The scale bar is 1 nm.

EDX mapping suggests that a combination of different defects accompanies dislocation formation that is driven by strain relaxation. It is, however, impractical to compute these defects simultaneously due to the size of the model and the large number of defect types and configurations that would have to be taken into account. We thus try to identify guiding rules for point defect formation in the vicinity of the dislocation core at the STO/LSAT interface by considering the most relevant defect types separately.
We start by investigating the formation of a single neutral Ta (VTa) and Al (VAl) vacancy, for which the experiment shows Ta atoms to diffuse into STO, while Al diffuses out of dislocation cores that appear elementally hollow. The map of relative formation energies in Figure 7a indicates that one of the most stable VTa configurations is at the STO/LSAT interface, closest to the dislocation core. Interestingly, all other VTa molecules have much larger formation energies (by 1.2 eV) and are thus much less favorable to form. More generally, we observe that VTa positions with low Ef values are located in a trapezoidal area below the dislocation core (see Figure 7a). This peculiar profile matches with the magnitude of the compressive strain in the LSAT region (Figure 2c), which is known to lower the formation of cation vacancies. (48) This result suggests that VTa strongly favors a trapezoidal area below the dislocation core, in agreement with the EDX maps and the STEM image contrast (Figure 2b). Instead, VAl formation is less sensitive to strain, consistent with the smaller radius of Al3+, favorable sites being both at the STO/LSAT interface close to the dislocation core as well as further into the LSAT substrate, in agreement with EDX mapping.

Figure 7

Figure 7. Color map of the relative (a) VTa and (b) VAl formation energy (ΔEf) around the dislocation core. The energy of the most stable defect is used as a reference in each plot. The shaded trapezoidal area in (a) indicates where VTa tends to form/segregate. Electronic DOS projected on the atoms at the dislocation core containing (c) one VAl and (d) one VTa, each at the most stable site at the STO/LSAT interface or (e) away from the interface in LSAT. (f–h) show charge density isosurfaces (10–2 e/Å3) in the energy range of the defect states highlighted in brown or orange in the DOS.

As expected, the formation of both VTa and VAl leads to the creation of holes, as shown by the empty peaks with an O character above the Fermi level in Figures 7c–e. These states are localized on O atoms at the dislocation core when the vacancy is formed close to the core or on O atoms adjacent to the defect when the vacancy is further from the core (Figures 7f–h). Filling of these defect states at the dislocation core by electron doping, for example, due to oxygen vacancies, could further favor VO formation at the dislocation core and result in conductivity along the dislocation line. For these reasons, we re-evaluated the formation of a VO in the presence of a VTa. The comparison of Figures 8a and 4a suggests that Ef for a VO is lowered due to the VTa, with a reduction of up to 3 eV for O positions close to the VTa and both in LSAT and STO. For neighboring VTa and VO at the dislocation core, the excess electrons due to VO formation partially heal the VTa hole state, as can be seen by comparing Figure 8b,c with Figure 7b,e.

Figure 8

Figure 8. (a) Color map of the relative VO formation energy (ΔEf) around the dislocation core in the presence of a VTa, indicated by the orange square. (b) DOS projected on the atoms at the dislocation core containing one VTa at the dislocation core and one VO at a neighboring site. (c) Charge density isosurfaces (10–2 e/Å3) in the energy range of the defect state highlighted orange in the DOS.

3.2.3. Substitutional Defects

Since the interdiffusion of B-site cations, namely, Ti, Al, and Ta, between LSAT and STO is clearly visible in the EDX maps (Figure 6), we also considered the possibility of Ti substituting Al (TiAl) or Ta (TiTa) in the LSAT substrate and of Ta substituting Ti (TaTi) in the STO film. When TaTi is formed in STO (Figure 9c), we observe fairly strong variations in the formation energy for the explored configurations, with low values at the interface adjacent to the dislocation core but also further up in STO, in line with EDX maps showing Ta diffusion to 3 or even more Ti layers from the interface. Results for Ti diffusion into LSAT, forming TiTa or TiAl (Figure 9a,b) are, instead, at odds with experiments and suggest that Ti should not be mainly located at the dislocation core, contrary to what is shown in Figure 6c.

Figure 9

Figure 9. Color map of the relative (a) TiAl, (b) TiTa, and (c) TaTi formation energy (ΔEf) around the dislocation core. The energy of the most stable defect is used in each plot as a reference.

This result suggests that the Ti/Ta/Al intermixing observed experimentally could be the result of a complex interplay between different defect types, some of which can lead to excess charges. To further investigate the effect of electron or hole doping on the cation interdiffusion, we select the two example defects, TaTi and VTa, that result in electron and hole doping, respectively, and reevaluate TiAl and TiTa formation in LSAT in the presence of these defects. Figure 10 reports the average Mulliken charges for Ti atoms in STO (Tibulk) or at the dislocation core (Ticore) for different single defects in the dislocation model. This data shows, indeed, that while Tibulk stays unaltered, Ticore is oxidized/reduced in the presence of a TaTi/VTa compared to the stoichiometric model. Instead, Ti-substitution in LSAT (TiTa or TiAl) does not significantly alter the charges of Tibulk or Ticore, but instead the substitutional Ti atom (Tisub) is oxidized compared to Tibulk, especially when substitution takes place at the Ta site. Comparison of Figures 9 and 11 indicates that reducing TaTi leading to electron doping at the dislocation core favors Ti-substitution around the dislocation at less oxidizing Al3+ sites, while hole doping by VTa at the dislocation core increases Ti-substitution at the more oxidizing Ta5+ site due to charge compensation reasons. In summary, these data show that while Ti diffusion into LSAT is unlikely without other defects, it is facilitated by the simultaneous presence of VTa or by Ta diffusion into STO (TaTi).

Figure 10

Figure 10. Average Mulliken charges for Ti atoms in STO far from the dislocation core (Tibulk), Ti atoms close to the dislocation core (Ticore), and substitutional Ti atoms in LSAT (Tisub).

Figure 11

Figure 11. Color map of the relative formation energy (ΔEf) of a TiAl in the presence of one (a) TaTi or (b) VTa and of a TiTa in the presence of one (c) TaTi or (d) VTa. The energy of the most stable defect is used in each plot as a reference. The black square indicates the position of the TaTi or VTa defect.

4. Conclusions

Click to copy section linkSection link copied!

In this work, we investigated the atomic structure and defect chemistry of the highly ordered Moiré network of dislocations formed at the interface between an STO thin film grown on an LSAT substrate, combining experimental and theoretical techniques. Combined MD and DFT calculations lead to an atomic-scale model of the dislocation core that features undercoordinated Ta/Al cations at the interface and has a strain field in nice agreement with the one derived from high-resolution STEM data.
Both EELS and DFT results indicate that oxygen vacancies (VO) easily form at the dislocation core and in the STO above the dislocation core. Furthermore, DFT calculations show that the experimentally observed Ti3+ around the dislocation core is due to both the dislocation structure itself and the presence of VO.
EDX mapping suggests that cation vacancies form at the dislocation core in LSAT: Ti substitutes Al and Ta around the dislocation core in LSAT, and a partial substitution of Ta by Ti takes place in the STO film above the dislocation core. DFT calculations confirm that cation vacancies are favored to form in a compressively strained region below the dislocation core, leading to hole doping that, in turn, further favors VO formation at the core. Finally, we show that Ti diffusion into the LSAT substrate below the dislocation core only occurs in the presence of cation vacancies (favoring substitutions at Ta sites) or concurrently with the diffusion of Ta into STO (favoring substitution at the Al site).
Even though additional defect combinations and sites further from the dislocation core could be explored, the present DFT results and their good agreement with the experiment lead to a deeper understanding of the structure and electronic properties of these systems. Our results show, in particular, the predominance of p-type 1D conductivity along the dislocations, depending on the defects present also with shallow acceptor states. These results will be instrumental in further engineering the functional properties of these systems.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Ulrich Aschauer - Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012 Bern, SwitzerlandDepartment of Chemistry and Physics of Materials, University of Salzburg, Jakob-Haringer-Street 2A, A-5020 Salzburg, AustriaOrcidhttps://orcid.org/0000-0002-1165-6377 Email: [email protected]
  • Authors
    • Chiara Ricca - Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012 Bern, Switzerland
    • Elizabeth Skoropata - Swiss Light Source, Paul Scherrer Institut, Forschungsstrasse 111, 5232 Villigen PSI, Switzerland
    • Marta D. Rossell - Electron Microscopy Center, Empa, Swiss Federal Laboratories for Materials Science and Technology, Überlandstrasse 129, 8600 Dübendorf, SwitzerlandOrcidhttps://orcid.org/0000-0001-8610-8853
    • Rolf Erni - Electron Microscopy Center, Empa, Swiss Federal Laboratories for Materials Science and Technology, Überlandstrasse 129, 8600 Dübendorf, SwitzerlandOrcidhttps://orcid.org/0000-0003-2391-5943
    • Urs Staub - Swiss Light Source, Paul Scherrer Institut, Forschungsstrasse 111, 5232 Villigen PSI, SwitzerlandOrcidhttps://orcid.org/0000-0003-2035-3367
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

C.R. and E.S. were supported by the NCCR MARVEL, a National Centre of Competence in Research, funded by the Swiss National Science Foundation (grant number 182892). E.S. was also supported by the European Union’s Horizon 2020 research and innovation program under the Marie Sklodowska-Curie grant agreement no 884104 (PSI-FELLOW-III-3i). Computational resources were provided by the University of Bern (on the HPC cluster UBELIX, http://www.id.unibe.ch/hpc) and by the Swiss National Supercomputing Center (CSCS) under project ID mr26.

References

Click to copy section linkSection link copied!

This article references 55 other publications.

  1. 1
    Zubko, P.; Gariglio, S.; Gabay, M.; Ghosez, P.; Triscone, J.-M. Interface Physics in Complex Oxide Heterostructures. Annu. Rev. Condens. Matter Phys. 2011, 2, 141165,  DOI: 10.1146/annurev-conmatphys-062910-140445
  2. 2
    Lin, X.; Zhu, Z.; Fauqué, B.; Behnia, K. Fermi Surface of the Most Dilute Superconductor. Phys. Rev. X 2013, 3, 021002,  DOI: 10.1103/PhysRevX.3.021002
  3. 3
    Bhattacharya, A.; May, S. J. Magnetic Oxide Heterostructures. Annu. Rev. Mater. Res. 2014, 44, 6590,  DOI: 10.1146/annurev-matsci-070813-113447
  4. 4
    Arandiyan, H.; S Mofarah, S.; Sorrell, C. C.; Doustkhah, E.; Sajjadi, B.; Hao, D.; Wang, Y.; Sun, H.; Ni, B.-J.; Rezaei, M.; Shao, Z.; Maschmeyer, T. Defect Engineering of Oxide Perovskites for Catalysis and Energy Storage: Synthesis of Chemistry and Materials Science. Chem. Soc. Rev. 2021, 50, 1011610211,  DOI: 10.1039/d0cs00639d
  5. 5
    Kalinin, S. V.; Spaldin, N. A. Functional Ion Defects in Transition Metal Oxides. Science 2013, 341, 858859,  DOI: 10.1126/science.1243098
  6. 6
    Chandrasena, R. U.; Yang, W.; Lei, Q.; Delgado-Jaime, M. U.; Wijesekara, K. D.; Golalikhani, M.; Davidson, B. A.; Arenholz, E.; Kobayashi, K.; Kobata, M.; de Groot, F. M. F.; Aschauer, U.; Spaldin, N. A.; Xi, X.; Gray, A. X. Strain-Engineered Oxygen Vacancies in CaMnO3 Thin Films. Nano Lett. 2017, 17, 794799,  DOI: 10.1021/acs.nanolett.6b03986
  7. 7
    Ricca, C.; Niederhauser, N.; Aschauer, U. Local Polarization in Oxygen-Deficient LaMnO3 Induced by Charge Localization in the Jahn-Teller Distorted Structure. Phys. Rev. Res. 2020, 2, 042040,  DOI: 10.1103/PhysRevResearch.2.042040
  8. 8
    Ohta, H.; Kim, S.; Mune, Y.; Mizoguchi, T.; Nomura, K.; Ohta, S.; Nomura, T.; Nakanishi, Y.; Ikuhara, Y.; Hirano, M.; Hosono, H.; Koumoto, K. Giant Thermoelectric Seebeck Coefficient of a Two-Dimensional Electron Gas in SrTiO3. Nat. Mater. 2007, 6, 129134,  DOI: 10.1038/nmat1821
  9. 9
    Garcia-Barriocanal, J.; Rivera-Calzada, A.; Varela, M.; Sefrioui, Z.; Iborra, E.; Leon, C.; Pennycook, S. J.; Santamaria, J. Colossal Ionic Conductivity at Interfaces of Epitaxial ZrO2:Y2O3/SrTiO3 Heterostructures. Science 2008, 321, 676680,  DOI: 10.1126/science.1156393
  10. 10
    Mannhart, J.; Schlom, D. G. Oxide Interfaces-An Opportunity for Electronics. Science 2010, 327, 16071611,  DOI: 10.1126/science.1181862
  11. 11
    Zubko, P.; Jecklin, N.; Torres-Pardo, A.; Aguado-Puente, P.; Gloter, A.; Lichtensteiger, C.; Junquera, J.; Stéphan, O.; Triscone, J.-M. Electrostatic Coupling and Local Structural Distortions at Interfaces in Ferroelectric/Paraelectric Superlattices. Nano Lett. 2012, 12, 28462851,  DOI: 10.1021/nl3003717
  12. 12
    Chen, Y.; Green, R. J.; Sutarto, R.; He, F.; Linderoth, S.; Sawatzky, G. A.; Pryds, N. Tuning the Two-Dimensional Electron Liquid at Oxide Interfaces by Buffer-Layer-Engineered Redox Reactions. Nano Lett. 2017, 17, 70627066,  DOI: 10.1021/acs.nanolett.7b03744
  13. 13
    Burian, M.; Pedrini, B. F.; Ortiz Hernandez, N.; Ueda, H.; Vaz, C. A. F.; Caputo, M.; Radovic, M.; Staub, U. Buried Moiré Supercells Through SrTiO3 Nanolayer Relaxation. Phys. Rev. Res. 2021, 3, 013225,  DOI: 10.1103/PhysRevResearch.3.013225
  14. 14
    Adler, S. B. Chemical Expansivity of Electrochemical Ceramics. J. Am. Ceram. Soc. 2004, 84, 21172119,  DOI: 10.1111/j.1151-2916.2001.tb00968.x
  15. 15
    Aschauer, U.; Pfenninger, R.; Selbach, S. M.; Grande, T.; Spaldin, N. A. Strain-Controlled Oxygen Vacancy Formation and Ordering in CaMnO3. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 88, 054111,  DOI: 10.1103/PhysRevB.88.054111
  16. 16
    Aidhy, D. S.; Liu, B.; Zhang, Y.; Weber, W. J. Strain-Induced Phase and Oxygen-Vacancy Stability in Ionic Interfaces From First-Principles Calculations. J. Phys. Chem. C 2014, 118, 3013930144,  DOI: 10.1021/jp507876m
  17. 17
    Yildiz, B. Stretching” the Energy Landscape of Oxides - Effects on Electrocatalysis and Diffusion. MRS Bull. 2014, 39, 147156,  DOI: 10.1557/mrs.2014.8
  18. 18
    Choudhury, S.; Morgan, D.; Uberuaga, B. P. Massive Interfacial Reconstruction at Misfit Dislocations in Metal/Oxide Interfaces. Sci. Rep. 2014, 4, 6533,  DOI: 10.1038/srep06533
  19. 19
    Dholabhai, P. P.; Uberuaga, B. P. Beyond Coherent Oxide Heterostructures: Atomic-Scale Structure of Misfit Dislocations. Adv. Theory Simul. 2019, 2, 1900078,  DOI: 10.1002/adts.201900078
  20. 20
    Marrocchelli, D.; Sun, L.; Yildiz, B. Dislocations in SrTiO3: Easy to Reduce but Not So Fast for Oxygen Transport. J. Am. Chem. Soc. 2015, 137, 47354748,  DOI: 10.1021/ja513176u
  21. 21
    Dholabhai, P. P.; Martinez, E.; Uberuaga, B. P. Influence of Chemistry and Misfit Dislocation Structure on Dopant Segregation at Complex Oxide Heterointerfaces. Adv. Theory Simul. 2019, 2, 1800095,  DOI: 10.1002/adts.201800095
  22. 22
    Marzano, C.; Dholabhai, P. P. High-Throughput Prediction of Thermodynamic Stabilities of Dopant-Defect Clusters at Misfit Dislocations in Perovskite Oxide Heterostructures. J. Phys. Chem. C 2023, 127, 1598815999,  DOI: 10.1021/acs.jpcc.3c02367
  23. 23
    Hirel, P.; Mrovec, M.; Elsässer, C. Atomistic Simulation Study of (110) Dislocations in Strontium Titanate. Acta Mater. 2012, 60, 329338,  DOI: 10.1016/j.actamat.2011.09.049
  24. 24
    Sun, L.; Marrocchelli, D.; Yildiz, B. Edge Dislocation Slows Down Oxide Ion Diffusion in Doped CeO2 by Segregation of Charged Defects. Nat. Commun. 2015, 6, 6294,  DOI: 10.1038/ncomms7294
  25. 25
    Ricca, C.; Skoropata, E.; Rossell, M. D.; Erni, R.; Staub, U.; Aschauer, U. Combined Theoretical and Experimental Study of the Moiré Dislocation Network at the SrTiO3-(La,Sr)(Al,Ta)O3 Interface. 2023, arXiv:2307.12572.
  26. 26
    Plimpton, S. Fast Parallel Algorithms for Short-Range Molecular Dynamics. J. Comput. Phys. 1995, 117, 119,  DOI: 10.1006/jcph.1995.1039
  27. 27
    Lewis, G. V.; Catlow, C. R. A. Potential Models for Ionic Oxides. J. Phys. C: Solid State Phys. 1985, 18, 11491161,  DOI: 10.1088/0022-3719/18/6/010
  28. 28
    Gale, J. D. GULP: A Computer Program for the Symmetry-Adapted Simulation of Solids. J. Chem. Soc., Faraday Trans. 1997, 93, 629637,  DOI: 10.1039/a606455h
  29. 29
    Gale, J. D.; Rohl, A. L. The General Utility Lattice Program (GULP). Mol. Simul. 2003, 29, 291341,  DOI: 10.1080/0892702031000104887
  30. 30
    Gale, J. D. GULP: Capabilities and Prospects. Z. für Kristallogr. - Cryst. Mater. 2005, 220, 552554,  DOI: 10.1524/zkri.220.5.552.65070
  31. 31
    Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid Metals. Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47, 558561,  DOI: 10.1103/PhysRevB.47.558
  32. 32
    Kresse, G.; Hafner, J. Ab Initio Molecular-Dynamics Simulation of the Liquid-Metal - Amorphous-Semiconductor Transition in Germanium. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 49, 1425114269,  DOI: 10.1103/PhysRevB.49.14251
  33. 33
    Kresse, G.; Furthmüller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 1550,  DOI: 10.1016/0927-0256(96)00008-0
  34. 34
    Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 1116911186,  DOI: 10.1103/physrevb.54.11169
  35. 35
    Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
  36. 36
    Blöchl, P. E. Projector Augmented-Wave Method. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 50, 1795317979,  DOI: 10.1103/PhysRevB.50.17953
  37. 37
    Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 59, 17581775,  DOI: 10.1103/PhysRevB.59.1758
  38. 38
    Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B: Solid State 1976, 13, 51885192,  DOI: 10.1103/PhysRevB.13.5188
  39. 39
    Kühne, T. D.; Iannuzzi, M.; Del Ben, M.; Rybkin, V. V.; Seewald, P.; Stein, F.; Laino, T.; Khaliullin, R. Z.; Schütt, O.; Schiffmann, F. CP2K: An Electronic Structure and Molecular Dynamics Software Package - Quickstep: Efficient and Accurate Electronic Structure Calculations. J. Chem. Phys. 2020, 152, 194103,  DOI: 10.1063/5.0007045
  40. 40
    Goedecker, S.; Teter, M.; Hutter, J. Separable Dual-Space Gaussian Pseudopotentials. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 17031710,  DOI: 10.1103/PhysRevB.54.1703
  41. 41
    VandeVondele, J.; Hutter, J. Gaussian Basis Sets for Accurate Calculations on Molecular Systems in Gas and Condensed Phases. J. Chem. Phys. 2007, 127, 114105,  DOI: 10.1063/1.2770708
  42. 42
    Freysoldt, C.; Grabowski, B.; Hickel, T.; Neugebauer, J.; Kresse, G.; Janotti, A.; Van de Walle, C. G. First-Principles Calculations for Point Defects in Solids. Rev. Mod. Phys. 2014, 86, 253305,  DOI: 10.1103/RevModPhys.86.253
  43. 43
    Schlepütz, C. M.; Mariager, S. O.; Pauli, S. A.; Feidenhans’l, R.; Willmott, P. R. Angle Calculations for a (2 + 3)-Type Diffractometer: Focus on Area Detectors. J. Appl. Crystallogr. 2011, 44, 7383,  DOI: 10.1107/S0021889810048922
  44. 44
    Kriegner, D.; Wintersberger, E.; Stangl, J. xrayutilities: A Versatile Tool for Reciprocal Space Conversion of Scattering Data Recorded With Linear and Area Detectors. J. Appl. Crystallogr. 2013, 46, 11621170,  DOI: 10.1107/S0021889813017214
  45. 45
    Ruiz Caridad, A.; Erni, R.; Vogel, A.; Rossell, M. D. Applications of a Novel Electron Energy Filter Combined With a Hybrid-Pixel Direct Electron Detector for the Analysis of Functional Oxides by STEM-EELS and Energy-Filtered Imaging. Micron 2022, 160, 103331,  DOI: 10.1016/j.micron.2022.103331
  46. 46
    Campanini, M.; Erni, R.; Yang, C.-H.; Ramesh, R.; Rossell, M. D. Periodic Giant Polarization Gradients in Doped BiFeO3 Thin Films. Nano Lett. 2018, 18, 717724,  DOI: 10.1021/acs.nanolett.7b03817
  47. 47
    Muller, D. A.; Nakagawa, N.; Ohtomo, A.; Grazul, J. L.; Hwang, H. Y. Atomic-Scale Imaging of Nanoengineered Oxygen Vacancy Profiles in SrTiO3. Nature 2004, 430, 657661,  DOI: 10.1038/nature02756
  48. 48
    Aschauer, U.; Vonrüti, N.; Spaldin, N. A. Effect of Epitaxial Strain on Cation and Anion Vacancy Formation in MnO. Phys. Rev. B: Condens. Matter Mater. Phys. 2015, 92, 054103,  DOI: 10.1103/PhysRevB.92.054103
  49. 49
    Abbate, M.; de Groot, F. M. F.; Fuggle, J. C.; Fujimori, A.; Tokura, Y.; Fujishima, Y.; Strebel, O.; Domke, M.; Kaindl, G.; van Elp, J.; Thole, B. T.; Sawatzky, G. A.; Sacchi, M.; Tsuda, N. Soft-X-Ray-Absorption Studies of the Location of Extra Charges Induced by Substitution in Controlled-Valence Materials. Phys. Rev. B: Condens. Matter Mater. Phys. 1991, 44, 54195422,  DOI: 10.1103/PhysRevB.44.5419
  50. 50
    Ricca, C.; Timrov, I.; Cococcioni, M.; Marzari, N.; Aschauer, U. Self-Consistent DFT+U+V Study of Oxygen Vacancies in SrTiO3. Phys. Rev. Res. 2020, 2, 023313,  DOI: 10.1103/PhysRevResearch.2.023313
  51. 51
    Ricci, D.; Bano, G.; Pacchioni, G.; Illas, F. Electronic Structure of a Neutral Oxygen Vacancy in SrTiO3. Phys. Rev. B: Condens. Matter Mater. Phys. 2003, 68, 224105,  DOI: 10.1103/PhysRevB.68.224105
  52. 52
    Evarestov, R. A.; Kotomin, E. A.; Zhukovskii, Y. F. DFT Study of a Single F Center in Cubic SrTiO3 Perovskite. Int. J. Quantum Chem. 2006, 106, 21732183,  DOI: 10.1002/qua.20855
  53. 53
    Zhukovskii, Y. F.; Kotomin, E. A.; Piskunov, S.; Ellis, D. A Comparative Ab Initio Study of Bulk and Surface Oxygen Vacancies in PbTiO3, PbZrO3 and SrTiO3 Perovskites. Solid State Commun. 2009, 149, 13591362,  DOI: 10.1016/j.ssc.2009.05.023
  54. 54
    Mitra, C.; Lin, C.; Robertson, J.; Demkov, A. A. Electronic Structure of Oxygen Vacancies in SrTiO3 and LaAlO3. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 86, 155105,  DOI: 10.1103/PhysRevB.86.155105
  55. 55
    Lin, C.; Mitra, C.; Demkov, A. A. Orbital Ordering Under Reduced Symmetry in Transition Metal Perovskites: Oxygen Vacancy in SrTiO3. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 86, 161102,  DOI: 10.1103/physrevb.86.161102

Cited By

Click to copy section linkSection link copied!

This article has not yet been cited by other publications.

ACS Applied Materials & Interfaces

Cite this: ACS Appl. Mater. Interfaces 2023, 15, 46, 53678–53687
Click to copy citationCitation copied!
https://doi.org/10.1021/acsami.3c10958
Published November 9, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

676

Altmetric

-

Citations

-
Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. (a) In-plane projection of the X-ray diffraction of the (22̅4) reflection of SrTiO3/LSAT thin films with a Moiré lattice. Line cuts through the (b) film (L = 4.000) and (c) substrate (L = 4.034) reflections show the periodicity of the Moiré lattice. (d) HAADF-STEM image of a STO thin film grown on a LSAT (001) substrate and (e) corresponding strain map along the x direction calculated by geometric phase analysis. The misfit dislocations appear as butterflylike features at the STO/LSAT interface with a compression region (in blue) and a tensile region (in yellow). (f) Structure of the periodically repeated relaxed MD model of the dislocation array at the SrTiO3/LSAT interface with the dislocations and the (2 × 107 × 60) simulation cell highlighted. Only the B–O bonding network and the A-site atoms are shown for clarity. Insets show the initial unrelaxed part of the structure with the removed STO plane and the extracted DFT cluster model, respectively. In the DFT model, the faded-out atoms are fixed during DFT geometry optimization to impose elastic boundary conditions.

    Figure 2

    Figure 2. (a) Map of the interatomic distances (in Å) for the A sublattice along the horizontal direction as obtained by PPA from the HAADF-STEM image in (b). (c) Map of horizontal interatomic A–A distances extracted from the DFT-relaxed dislocation core structure shown in (d). Circles are located at the midpoint of A–A pairs and are color-coded according to the distance between that pair. (e) Electronic density of states (DOS) projected on the atoms at the dislocation core for the stoichiometric dislocation model. The vertical dashed line indicates the position of the Fermi level. (f) Charge density isosurface (10–2 e/Å3) in the energy range highlighted in red in (e).

    Figure 3

    Figure 3. (a) HAADF-STEM image of a dislocation at the STO/LSAT interface and corresponding atomic-resolution maps of the (b) O–K and (c) Ti–L3,2 excitation edges. The scale bar is 1 nm. (d) Representative Ti–L3,2 EELS spectra of the STO film and the surroundings of the dislocation core. The spectra have been normalized to the L2 eg peak height for clarity.

    Figure 4

    Figure 4. (a) Color map of the relative VO formation energy (ΔEf) around the dislocation core. Electronic DOS projected on the atoms at the dislocation core for the STO/LSAT interface model containing (b) one, (c) two, and (d) three VO in the most stable configurations. The vertical dashed line indicates the position of the Fermi level.

    Figure 5

    Figure 5. Mulliken charges for the Ti atoms in (a) the stoichiometric dislocation model and in the presence of (b) one, (c) two, and (d) three VO. The black squares indicate the position of the VO.

    Figure 6

    Figure 6. (a) HAADF-STEM image of a misfit dislocation at the STO/LSAT interface and (b–g) corresponding elemental maps of Sr, Ti, O, La, Al, and Ta calculated from an EDX spectrum image using the Sr–K, Ti–K, O–K1, La–L, Al–K1, and Ta–L lines, respectively. The scale bar is 1 nm.

    Figure 7

    Figure 7. Color map of the relative (a) VTa and (b) VAl formation energy (ΔEf) around the dislocation core. The energy of the most stable defect is used as a reference in each plot. The shaded trapezoidal area in (a) indicates where VTa tends to form/segregate. Electronic DOS projected on the atoms at the dislocation core containing (c) one VAl and (d) one VTa, each at the most stable site at the STO/LSAT interface or (e) away from the interface in LSAT. (f–h) show charge density isosurfaces (10–2 e/Å3) in the energy range of the defect states highlighted in brown or orange in the DOS.

    Figure 8

    Figure 8. (a) Color map of the relative VO formation energy (ΔEf) around the dislocation core in the presence of a VTa, indicated by the orange square. (b) DOS projected on the atoms at the dislocation core containing one VTa at the dislocation core and one VO at a neighboring site. (c) Charge density isosurfaces (10–2 e/Å3) in the energy range of the defect state highlighted orange in the DOS.

    Figure 9

    Figure 9. Color map of the relative (a) TiAl, (b) TiTa, and (c) TaTi formation energy (ΔEf) around the dislocation core. The energy of the most stable defect is used in each plot as a reference.

    Figure 10

    Figure 10. Average Mulliken charges for Ti atoms in STO far from the dislocation core (Tibulk), Ti atoms close to the dislocation core (Ticore), and substitutional Ti atoms in LSAT (Tisub).

    Figure 11

    Figure 11. Color map of the relative formation energy (ΔEf) of a TiAl in the presence of one (a) TaTi or (b) VTa and of a TiTa in the presence of one (c) TaTi or (d) VTa. The energy of the most stable defect is used in each plot as a reference. The black square indicates the position of the TaTi or VTa defect.

  • References


    This article references 55 other publications.

    1. 1
      Zubko, P.; Gariglio, S.; Gabay, M.; Ghosez, P.; Triscone, J.-M. Interface Physics in Complex Oxide Heterostructures. Annu. Rev. Condens. Matter Phys. 2011, 2, 141165,  DOI: 10.1146/annurev-conmatphys-062910-140445
    2. 2
      Lin, X.; Zhu, Z.; Fauqué, B.; Behnia, K. Fermi Surface of the Most Dilute Superconductor. Phys. Rev. X 2013, 3, 021002,  DOI: 10.1103/PhysRevX.3.021002
    3. 3
      Bhattacharya, A.; May, S. J. Magnetic Oxide Heterostructures. Annu. Rev. Mater. Res. 2014, 44, 6590,  DOI: 10.1146/annurev-matsci-070813-113447
    4. 4
      Arandiyan, H.; S Mofarah, S.; Sorrell, C. C.; Doustkhah, E.; Sajjadi, B.; Hao, D.; Wang, Y.; Sun, H.; Ni, B.-J.; Rezaei, M.; Shao, Z.; Maschmeyer, T. Defect Engineering of Oxide Perovskites for Catalysis and Energy Storage: Synthesis of Chemistry and Materials Science. Chem. Soc. Rev. 2021, 50, 1011610211,  DOI: 10.1039/d0cs00639d
    5. 5
      Kalinin, S. V.; Spaldin, N. A. Functional Ion Defects in Transition Metal Oxides. Science 2013, 341, 858859,  DOI: 10.1126/science.1243098
    6. 6
      Chandrasena, R. U.; Yang, W.; Lei, Q.; Delgado-Jaime, M. U.; Wijesekara, K. D.; Golalikhani, M.; Davidson, B. A.; Arenholz, E.; Kobayashi, K.; Kobata, M.; de Groot, F. M. F.; Aschauer, U.; Spaldin, N. A.; Xi, X.; Gray, A. X. Strain-Engineered Oxygen Vacancies in CaMnO3 Thin Films. Nano Lett. 2017, 17, 794799,  DOI: 10.1021/acs.nanolett.6b03986
    7. 7
      Ricca, C.; Niederhauser, N.; Aschauer, U. Local Polarization in Oxygen-Deficient LaMnO3 Induced by Charge Localization in the Jahn-Teller Distorted Structure. Phys. Rev. Res. 2020, 2, 042040,  DOI: 10.1103/PhysRevResearch.2.042040
    8. 8
      Ohta, H.; Kim, S.; Mune, Y.; Mizoguchi, T.; Nomura, K.; Ohta, S.; Nomura, T.; Nakanishi, Y.; Ikuhara, Y.; Hirano, M.; Hosono, H.; Koumoto, K. Giant Thermoelectric Seebeck Coefficient of a Two-Dimensional Electron Gas in SrTiO3. Nat. Mater. 2007, 6, 129134,  DOI: 10.1038/nmat1821
    9. 9
      Garcia-Barriocanal, J.; Rivera-Calzada, A.; Varela, M.; Sefrioui, Z.; Iborra, E.; Leon, C.; Pennycook, S. J.; Santamaria, J. Colossal Ionic Conductivity at Interfaces of Epitaxial ZrO2:Y2O3/SrTiO3 Heterostructures. Science 2008, 321, 676680,  DOI: 10.1126/science.1156393
    10. 10
      Mannhart, J.; Schlom, D. G. Oxide Interfaces-An Opportunity for Electronics. Science 2010, 327, 16071611,  DOI: 10.1126/science.1181862
    11. 11
      Zubko, P.; Jecklin, N.; Torres-Pardo, A.; Aguado-Puente, P.; Gloter, A.; Lichtensteiger, C.; Junquera, J.; Stéphan, O.; Triscone, J.-M. Electrostatic Coupling and Local Structural Distortions at Interfaces in Ferroelectric/Paraelectric Superlattices. Nano Lett. 2012, 12, 28462851,  DOI: 10.1021/nl3003717
    12. 12
      Chen, Y.; Green, R. J.; Sutarto, R.; He, F.; Linderoth, S.; Sawatzky, G. A.; Pryds, N. Tuning the Two-Dimensional Electron Liquid at Oxide Interfaces by Buffer-Layer-Engineered Redox Reactions. Nano Lett. 2017, 17, 70627066,  DOI: 10.1021/acs.nanolett.7b03744
    13. 13
      Burian, M.; Pedrini, B. F.; Ortiz Hernandez, N.; Ueda, H.; Vaz, C. A. F.; Caputo, M.; Radovic, M.; Staub, U. Buried Moiré Supercells Through SrTiO3 Nanolayer Relaxation. Phys. Rev. Res. 2021, 3, 013225,  DOI: 10.1103/PhysRevResearch.3.013225
    14. 14
      Adler, S. B. Chemical Expansivity of Electrochemical Ceramics. J. Am. Ceram. Soc. 2004, 84, 21172119,  DOI: 10.1111/j.1151-2916.2001.tb00968.x
    15. 15
      Aschauer, U.; Pfenninger, R.; Selbach, S. M.; Grande, T.; Spaldin, N. A. Strain-Controlled Oxygen Vacancy Formation and Ordering in CaMnO3. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 88, 054111,  DOI: 10.1103/PhysRevB.88.054111
    16. 16
      Aidhy, D. S.; Liu, B.; Zhang, Y.; Weber, W. J. Strain-Induced Phase and Oxygen-Vacancy Stability in Ionic Interfaces From First-Principles Calculations. J. Phys. Chem. C 2014, 118, 3013930144,  DOI: 10.1021/jp507876m
    17. 17
      Yildiz, B. Stretching” the Energy Landscape of Oxides - Effects on Electrocatalysis and Diffusion. MRS Bull. 2014, 39, 147156,  DOI: 10.1557/mrs.2014.8
    18. 18
      Choudhury, S.; Morgan, D.; Uberuaga, B. P. Massive Interfacial Reconstruction at Misfit Dislocations in Metal/Oxide Interfaces. Sci. Rep. 2014, 4, 6533,  DOI: 10.1038/srep06533
    19. 19
      Dholabhai, P. P.; Uberuaga, B. P. Beyond Coherent Oxide Heterostructures: Atomic-Scale Structure of Misfit Dislocations. Adv. Theory Simul. 2019, 2, 1900078,  DOI: 10.1002/adts.201900078
    20. 20
      Marrocchelli, D.; Sun, L.; Yildiz, B. Dislocations in SrTiO3: Easy to Reduce but Not So Fast for Oxygen Transport. J. Am. Chem. Soc. 2015, 137, 47354748,  DOI: 10.1021/ja513176u
    21. 21
      Dholabhai, P. P.; Martinez, E.; Uberuaga, B. P. Influence of Chemistry and Misfit Dislocation Structure on Dopant Segregation at Complex Oxide Heterointerfaces. Adv. Theory Simul. 2019, 2, 1800095,  DOI: 10.1002/adts.201800095
    22. 22
      Marzano, C.; Dholabhai, P. P. High-Throughput Prediction of Thermodynamic Stabilities of Dopant-Defect Clusters at Misfit Dislocations in Perovskite Oxide Heterostructures. J. Phys. Chem. C 2023, 127, 1598815999,  DOI: 10.1021/acs.jpcc.3c02367
    23. 23
      Hirel, P.; Mrovec, M.; Elsässer, C. Atomistic Simulation Study of (110) Dislocations in Strontium Titanate. Acta Mater. 2012, 60, 329338,  DOI: 10.1016/j.actamat.2011.09.049
    24. 24
      Sun, L.; Marrocchelli, D.; Yildiz, B. Edge Dislocation Slows Down Oxide Ion Diffusion in Doped CeO2 by Segregation of Charged Defects. Nat. Commun. 2015, 6, 6294,  DOI: 10.1038/ncomms7294
    25. 25
      Ricca, C.; Skoropata, E.; Rossell, M. D.; Erni, R.; Staub, U.; Aschauer, U. Combined Theoretical and Experimental Study of the Moiré Dislocation Network at the SrTiO3-(La,Sr)(Al,Ta)O3 Interface. 2023, arXiv:2307.12572.
    26. 26
      Plimpton, S. Fast Parallel Algorithms for Short-Range Molecular Dynamics. J. Comput. Phys. 1995, 117, 119,  DOI: 10.1006/jcph.1995.1039
    27. 27
      Lewis, G. V.; Catlow, C. R. A. Potential Models for Ionic Oxides. J. Phys. C: Solid State Phys. 1985, 18, 11491161,  DOI: 10.1088/0022-3719/18/6/010
    28. 28
      Gale, J. D. GULP: A Computer Program for the Symmetry-Adapted Simulation of Solids. J. Chem. Soc., Faraday Trans. 1997, 93, 629637,  DOI: 10.1039/a606455h
    29. 29
      Gale, J. D.; Rohl, A. L. The General Utility Lattice Program (GULP). Mol. Simul. 2003, 29, 291341,  DOI: 10.1080/0892702031000104887
    30. 30
      Gale, J. D. GULP: Capabilities and Prospects. Z. für Kristallogr. - Cryst. Mater. 2005, 220, 552554,  DOI: 10.1524/zkri.220.5.552.65070
    31. 31
      Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid Metals. Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47, 558561,  DOI: 10.1103/PhysRevB.47.558
    32. 32
      Kresse, G.; Hafner, J. Ab Initio Molecular-Dynamics Simulation of the Liquid-Metal - Amorphous-Semiconductor Transition in Germanium. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 49, 1425114269,  DOI: 10.1103/PhysRevB.49.14251
    33. 33
      Kresse, G.; Furthmüller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 1550,  DOI: 10.1016/0927-0256(96)00008-0
    34. 34
      Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 1116911186,  DOI: 10.1103/physrevb.54.11169
    35. 35
      Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
    36. 36
      Blöchl, P. E. Projector Augmented-Wave Method. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 50, 1795317979,  DOI: 10.1103/PhysRevB.50.17953
    37. 37
      Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 59, 17581775,  DOI: 10.1103/PhysRevB.59.1758
    38. 38
      Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B: Solid State 1976, 13, 51885192,  DOI: 10.1103/PhysRevB.13.5188
    39. 39
      Kühne, T. D.; Iannuzzi, M.; Del Ben, M.; Rybkin, V. V.; Seewald, P.; Stein, F.; Laino, T.; Khaliullin, R. Z.; Schütt, O.; Schiffmann, F. CP2K: An Electronic Structure and Molecular Dynamics Software Package - Quickstep: Efficient and Accurate Electronic Structure Calculations. J. Chem. Phys. 2020, 152, 194103,  DOI: 10.1063/5.0007045
    40. 40
      Goedecker, S.; Teter, M.; Hutter, J. Separable Dual-Space Gaussian Pseudopotentials. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 17031710,  DOI: 10.1103/PhysRevB.54.1703
    41. 41
      VandeVondele, J.; Hutter, J. Gaussian Basis Sets for Accurate Calculations on Molecular Systems in Gas and Condensed Phases. J. Chem. Phys. 2007, 127, 114105,  DOI: 10.1063/1.2770708
    42. 42
      Freysoldt, C.; Grabowski, B.; Hickel, T.; Neugebauer, J.; Kresse, G.; Janotti, A.; Van de Walle, C. G. First-Principles Calculations for Point Defects in Solids. Rev. Mod. Phys. 2014, 86, 253305,  DOI: 10.1103/RevModPhys.86.253
    43. 43
      Schlepütz, C. M.; Mariager, S. O.; Pauli, S. A.; Feidenhans’l, R.; Willmott, P. R. Angle Calculations for a (2 + 3)-Type Diffractometer: Focus on Area Detectors. J. Appl. Crystallogr. 2011, 44, 7383,  DOI: 10.1107/S0021889810048922
    44. 44
      Kriegner, D.; Wintersberger, E.; Stangl, J. xrayutilities: A Versatile Tool for Reciprocal Space Conversion of Scattering Data Recorded With Linear and Area Detectors. J. Appl. Crystallogr. 2013, 46, 11621170,  DOI: 10.1107/S0021889813017214
    45. 45
      Ruiz Caridad, A.; Erni, R.; Vogel, A.; Rossell, M. D. Applications of a Novel Electron Energy Filter Combined With a Hybrid-Pixel Direct Electron Detector for the Analysis of Functional Oxides by STEM-EELS and Energy-Filtered Imaging. Micron 2022, 160, 103331,  DOI: 10.1016/j.micron.2022.103331
    46. 46
      Campanini, M.; Erni, R.; Yang, C.-H.; Ramesh, R.; Rossell, M. D. Periodic Giant Polarization Gradients in Doped BiFeO3 Thin Films. Nano Lett. 2018, 18, 717724,  DOI: 10.1021/acs.nanolett.7b03817
    47. 47
      Muller, D. A.; Nakagawa, N.; Ohtomo, A.; Grazul, J. L.; Hwang, H. Y. Atomic-Scale Imaging of Nanoengineered Oxygen Vacancy Profiles in SrTiO3. Nature 2004, 430, 657661,  DOI: 10.1038/nature02756
    48. 48
      Aschauer, U.; Vonrüti, N.; Spaldin, N. A. Effect of Epitaxial Strain on Cation and Anion Vacancy Formation in MnO. Phys. Rev. B: Condens. Matter Mater. Phys. 2015, 92, 054103,  DOI: 10.1103/PhysRevB.92.054103
    49. 49
      Abbate, M.; de Groot, F. M. F.; Fuggle, J. C.; Fujimori, A.; Tokura, Y.; Fujishima, Y.; Strebel, O.; Domke, M.; Kaindl, G.; van Elp, J.; Thole, B. T.; Sawatzky, G. A.; Sacchi, M.; Tsuda, N. Soft-X-Ray-Absorption Studies of the Location of Extra Charges Induced by Substitution in Controlled-Valence Materials. Phys. Rev. B: Condens. Matter Mater. Phys. 1991, 44, 54195422,  DOI: 10.1103/PhysRevB.44.5419
    50. 50
      Ricca, C.; Timrov, I.; Cococcioni, M.; Marzari, N.; Aschauer, U. Self-Consistent DFT+U+V Study of Oxygen Vacancies in SrTiO3. Phys. Rev. Res. 2020, 2, 023313,  DOI: 10.1103/PhysRevResearch.2.023313
    51. 51
      Ricci, D.; Bano, G.; Pacchioni, G.; Illas, F. Electronic Structure of a Neutral Oxygen Vacancy in SrTiO3. Phys. Rev. B: Condens. Matter Mater. Phys. 2003, 68, 224105,  DOI: 10.1103/PhysRevB.68.224105
    52. 52
      Evarestov, R. A.; Kotomin, E. A.; Zhukovskii, Y. F. DFT Study of a Single F Center in Cubic SrTiO3 Perovskite. Int. J. Quantum Chem. 2006, 106, 21732183,  DOI: 10.1002/qua.20855
    53. 53
      Zhukovskii, Y. F.; Kotomin, E. A.; Piskunov, S.; Ellis, D. A Comparative Ab Initio Study of Bulk and Surface Oxygen Vacancies in PbTiO3, PbZrO3 and SrTiO3 Perovskites. Solid State Commun. 2009, 149, 13591362,  DOI: 10.1016/j.ssc.2009.05.023
    54. 54
      Mitra, C.; Lin, C.; Robertson, J.; Demkov, A. A. Electronic Structure of Oxygen Vacancies in SrTiO3 and LaAlO3. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 86, 155105,  DOI: 10.1103/PhysRevB.86.155105
    55. 55
      Lin, C.; Mitra, C.; Demkov, A. A. Orbital Ordering Under Reduced Symmetry in Transition Metal Perovskites: Oxygen Vacancy in SrTiO3. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 86, 161102,  DOI: 10.1103/physrevb.86.161102