In polar crystals, cooperative ionic displacement produces a macroscopic spontaneous polarization. Among such polar materials, LiNbO3-type wide bandgap oxides are particularly appealing because they offer useful ferroelectric properties and also potentially lead to multiferroic materials. Using molecular-beam epitaxy, we investigated the thin-film growth of high-pressure phase LiNbO3-type ZnSnO3 and discovered a polar oxide candidate, MgSnO3. We found that LiNbO3-type substrates play an essential role in the crystallization of these compounds, though corundum-type Al2O3 substrates also have the identical crystallographic arrangement of oxygen sublattice. Optical transmittance and electrical transport measurements revealed their potential as a transparent conducting oxide. Establishment of a thin-film synthetic route would be the basis for exploration of functional polar oxides and research on conduction at ferroelectric interfaces and domain walls.

Awide spectrum of functional properties exhibited by polar crystals1 are at the heart of many devices including pyroelectric sensors, optical wavelength converters, and ferroelectric memories.2 One of the representative polar systems is perovskite-derivative ABO3; the ideal perovskite structure is cubic and nonpolar. The lattice distortion from cubic perovskite and appearance of polar phases are guided with the tolerance factor,3,4t, given by t=rA+rO/2rB+rO, where rA, rB, and rO are the ionic radii of the A-site cation, B-site cation, and oxygen ion, respectively. The Goldschmidt diagram,4 presented in Fig. 1(a), indicates that the deviation of t from unity (ideal cubic structure) involves the BO6 octahedral tilting. In the presence of second-order Jahn-Teller active ions or lone-pair ions, the cubic instability increases and, at low temperature, often gives rise to a structural transformation to a low-symmetry polar phase. Archetypal ferroelectric oxides like PbTiO3 and BiFeO3 are classified into this category. In the low-t region (t < 0.8), corundum-derivative ABO3 structures become predominant, known as LiNbO3-type and ilmenite-type. The LiNbO3-type structure is polar (space group: R3c), while the ilmenite-type one is nonpolar (R3¯). Recently, a series of new LiNbO3-type polar ABO3 such as ZnSnO3, MnSnO3, and FeTiO3 have been synthesized by high-pressure techniques.5–11 In those oxides, in addition to ferroelectric polarization as large as those of LiNbO3 and LiTaO3,12 control of electronic and magnetic properties by utilizing various combinations of two cations is anticipated. Couplings of charge and spin degrees of freedom via the low symmetry character would also lead to candidates for multiferroic materials.6,10,11,13,14

FIG. 1.

(a) Goldschmidt diagram showing the structural stability of perovskite-derivative (represented by squares) and corundum-derivative (circle and triangles) phases as functions of cation radii rA and rB. See text for the definition of tolerance factor, t. Schematic illustrations in the inset (drawn with VESTA24) depict the evolution of BO6 octahedral tilting. (b) Topmost surface atomic arrangement in three-fold symmetry planes of typical oxide substrates, SrTiO3(111), Al2O3(0001), and LiNbO3(0001). The gray colored octahedra are located beneath the topmost layer. Cross-sectional views along the dashed lines are shown in the lower panels for Al2O3(0001) and LiNbO3(0001) (one unit cells) (Ref. 4). The spheres represent A and B cations accommodated in the oxygen octahedra, illustrating the LiNbO3-type atomic ordering.

FIG. 1.

(a) Goldschmidt diagram showing the structural stability of perovskite-derivative (represented by squares) and corundum-derivative (circle and triangles) phases as functions of cation radii rA and rB. See text for the definition of tolerance factor, t. Schematic illustrations in the inset (drawn with VESTA24) depict the evolution of BO6 octahedral tilting. (b) Topmost surface atomic arrangement in three-fold symmetry planes of typical oxide substrates, SrTiO3(111), Al2O3(0001), and LiNbO3(0001). The gray colored octahedra are located beneath the topmost layer. Cross-sectional views along the dashed lines are shown in the lower panels for Al2O3(0001) and LiNbO3(0001) (one unit cells) (Ref. 4). The spheres represent A and B cations accommodated in the oxygen octahedra, illustrating the LiNbO3-type atomic ordering.

Close modal

One important step toward further exploration of electronic functionality in relevant LiNbO3-type oxides is the thin-film growth.13–15 Oxide film growth generally begins with considerations of oxygen sublattice matching at the film/substrate interface.16 The averaged oxygen-oxygen distance and oxygen sublattice symmetry are compared between substrate and overgrown film materials. In Fig. 1(b), three-fold symmetry planes of typical oxide substrates, SrTiO3(111), Al2O3(0001), and LiNbO3(0001), are depicted. For simplicity, only the topmost oxygen ions in AO6 and BO6 octahedra are shown by small red spheres. These crystal planes have similar oxygen frameworks and are suited to thin-film growth of oxides with triangular lattices. In particular, corundum-type A2O3 and LiNbO3-type ABO3 have identical arrangement of oxygen sublattice and octahedral vacancy sites. As seen in the schematic cross sections, however, periodic cation occupation is inherent to the LiNbO3-type structure. In the film growth of (111)-oriented cubic perovskite-type oxides, such a cation geometry has not been considered much. One of the reasons for this may be that their large A and small B cations are coordinated by twelve and eight oxygen ions, respectively, and thus hardly occupy the anti-sites. In the case for LiNbO3-type oxides, the situation is however not so clear; the similar octahedral structural units as well as comparable ionic radii might give rise to disordered cation arrangement [Fig. 1(b)]. By molecular-beam epitaxy on LiNbO3(0001) substrates, we recently succeeded in stabilizing the LiNbO3-type ZnSnO3 film and demonstrated its high-mobility transistor operation.17 In this paper, we describe the decisive role of substrate in the crystallization and the thin-film synthesis of a new LiNbO3-type oxide, MgSnO3. These polar stannates not only are transparent oxide semiconductors exhibiting good electrical conduction but also can be a candidate for exploring a polar metal phase.

The films were grown on SrTiO3(111), Al2O3(0001), LiNbO3(0001), and LiNbO3-type LiTaO3(0001) substrates by oxygen-plasma assisted molecular-beam epitaxy. Two-side polished, pre-poled, congruent LiNbO3 and LiTaO3 single crystal substrates were mainly used, and stoichiometric LiNbO3 substrates were also used for scanning transmission electron microscopy (STEM). Prior to the film growth, SrTiO3, Al2O3, congruent LiNbO3, and stoichiometric LiNbO3 substrates were annealed in an oxygen ambient at 900 °C, 1000 °C, 900 °C, and 950 °C, respectively. LiTaO3 was not annealed to keep the pre-poled state because the ferroelectric transition temperature is as low as 600 °C. Elemental Sn, Zn, Mg, and Mn were supplied from effusion cells with high-purity metal sources. Beam equivalent pressure was monitored with a nude ion gauge and was controlled by the cell temperature. Typical pressures were 5.0 × 10−6 Pa, 1.1 × 10−5 Pa, 3.5 × 10−6 Pa, and 1.1 × 10−5 Pa for Sn, Zn, Mg, and Mn, respectively. Atomic oxygen radicals were generated from O2 gas using a radio frequency plasma cell. All the films presented below were fabricated at a substrate temperature of 500 °C. The Zn/Sn, Mg/Sn, and Mn/Sn atomic ratios in the films were adjusted to be ∼1 by compositional analysis with X-ray photoelectron spectroscopy and energy dispersive X-ray spectroscopy (see Fig. S1 and Methods in the supplementary material).

In Fig. 2(a), out-of-plane X-ray diffraction (XRD) patterns for Zn-Sn-O films on LiNbO3 (negatively poled −Z surface), Al2O3, and SrTiO3 substrates with Zn/Sn atomic ratios close to unity are compared. An intense ZnSnO3(0006) reflection with thickness fringes is observed for the film grown on LiNbO3(0001). The c-axis length is the closest to the bulk value of c = 14.00 Å around the nearly stoichiometric condition (Fig. S2), verifying the validity of our scheme for growth condition optimization. The film surface has a step-and-terrace structure [inset of Fig. 2(a)], and the step height is approximately one sixth of the c-axis length of ZnSnO3 [Fig. 1(b)]. The c-axis oriented ZnSnO3 film can also be grown on the positively poled + Z surface of LiNbO3 [the top panel in Fig. 2(b)] and isostructural LiTaO3(0001) (Fig. S3). In stark contrast, no traces of crystalline phases are discerned on Al2O3(0001) and SrTiO3(111). Note that this distinct growth behavior was not merely due to substrate-dependent optimal growth conditions; even in the examinations of the Zn/Sn flux ratio and substrate temperature, the film growth was not successful.

FIG. 2.

(a) Out-of-plane XRD patterns for Zn-Sn-O films grown on congruent LiNbO3(0001) (−Z surface), Al2O3(0001), and SrTiO3(111) with a Zn/Sn flux ratio of 2.2. (b) Data for a ZnSnO3 film on congruent LiNbO3(0001) (+Z surface) and MgSnO3 and MnSnO3 films on congruent LiNbO3(0001) (−Z surface). The film thicknesses were approximately 100 nm. The insets display typical atomic force microscopy images.

FIG. 2.

(a) Out-of-plane XRD patterns for Zn-Sn-O films grown on congruent LiNbO3(0001) (−Z surface), Al2O3(0001), and SrTiO3(111) with a Zn/Sn flux ratio of 2.2. (b) Data for a ZnSnO3 film on congruent LiNbO3(0001) (+Z surface) and MgSnO3 and MnSnO3 films on congruent LiNbO3(0001) (−Z surface). The film thicknesses were approximately 100 nm. The insets display typical atomic force microscopy images.

Close modal

We examined the effectiveness of LiNbO3-type substrates on the growth of other polar stannate candidates, MgSnO3 and MnSnO3. While LiNbO3-type MnSnO3 is synthesized by a high-pressure technique,6 little has been known about structural variants of MgSnO3; only the ilmenite-type form is known experimentally.18,19 As shown in the middle panel in Fig. 2(b), a highly crystallized phase is obtained on LiNbO3(0001) with a Mg/Sn flux ratio of 0.7. The XRD pattern satisfies the extinction rule for the LiNbO3-type structure (see Fig. S4 for the wide scan), the peak angle of which is close to that of ZnSnO3. A characteristic step-and-terrace structure, suggestive of common growth behavior, is also observed (inset). More importantly, the crystallization again proceeds only on LiNbO3-type substrates but not on Al2O3 or SrTiO3 substrates (Figs. S3 and S5). The Mn–Sn–O film also shows a weak but clear peak, which is consistent with the high-pressure LiNbO3-type MnSnO3 phase6 [the bottom panel in Fig. 2(b)].

On the (0001) plane of LiNbO3-type crystals where two distinct cations and metal vacancy exist, the oxygen-oxygen distance LO-O consists of three different values of 3.358, 2.884, and 2.723 Å for LiNbO3, and 3.334, 3.022, and 2.784 Å for ZnSnO3. By contrast, there are two LO-O (2.862 and 2.523 Å with a ratio of 2:1) on corundum-type Al2O3(0001) and only one LO-O (2.761 Å) on perovskite-type SrTiO3(111). Using the averaged LO-O value, the lattice mismatch between ZnSnO3(0001) and the substrate is calculated to be 2.0%, 10.8%, and 10.4% for LiNbO3(0001), Al2O3(0001), and SrTiO3(111), respectively. In preceding studies using pulsed-laser deposition, it was reported that highly crystalline films of high-pressure phase NiTiO313,14 and ZnSnO315 as well as ambient phase LiNbO3 (Refs. 20 and 21) can be obtained on Al2O3(0001) and SrTiO3(111) despite of the large lattice mismatch. Therefore, the observed substrate dependent growth points to the different growth mechanism of high-pressure phase stannates by molecular-beam epitaxy. We speculate that periodic modulation of LO-O and surface potential might act as a seed for the LiNbO3-type order.

Lattice parameters were evaluated for the highly crystalline stannate films by XRD reciprocal space mapping (RSM).Figure 3(a) displays RSM around LiNbO3(112¯12̲). This ZnSnO3 film on LiNbO3(0001) (−Z surface) is the same to that used for the out-of-plane XRD measurement [Fig. 2(a)]. The film is partially relaxed, and the lattice parameters are calculated to be a = 5.236 Å and c = 14.25 Å, which are comparable to a = 5.262 Å and c = 14.00 Å reported for bulk.5 A locally modified ilmenite-like atomic arrangement at the domain boundaries17 is probably the origin of the c-axis elongation of 2% from the bulk value of ZnSnO3. A similar result is obtained for the Mg–Sn–O film [Fig. 2(b)], shown in Fig. 3(b), yielding a = 5.211 Å and c = 14.24 Å. The shrinkage of lattice parameters as compared to those of ZnSnO3 is consistent with the smaller ionic radius of Mg2+ (0.72 Å) than that of Zn2+ (0.74 Å), which supports the introduction of Mg into the A-site. Based on these results, we conclude that LiNbO3-type MgSnO3 is stabilized in the thin-film form. From the inspection of in-plane rotational symmetry by ϕ-scan measurement [Fig. 3(c)], these c-axis oriented films were found to contain two domains of three-fold symmetry, rotated by 60° in the (0001) plane.

FIG. 3.

RSM around (112¯12̲) for (a) ZnSnO3 and (b) MgSnO3 films on congruent LiNbO3(0001) (−Z surface). These films were the same to those used for the out-of-plane XRD measurement shown in Fig. 2. (c) XRD ϕ-scan around (101¯4) measured for 160-nm-thick ZnSnO3 and 180-nm-thick MgSnO3 films on congruent LiNbO3(0001) (−Z surface) and a bare LiNbO3 substrate.

FIG. 3.

RSM around (112¯12̲) for (a) ZnSnO3 and (b) MgSnO3 films on congruent LiNbO3(0001) (−Z surface). These films were the same to those used for the out-of-plane XRD measurement shown in Fig. 2. (c) XRD ϕ-scan around (101¯4) measured for 160-nm-thick ZnSnO3 and 180-nm-thick MgSnO3 films on congruent LiNbO3(0001) (−Z surface) and a bare LiNbO3 substrate.

Close modal

Atomically resolved STEM imaging gives further evidence for the formation of LiNbO3-type ZnSnO3.Figure 4(a) shows a high angle annular dark field (HAADF) STEM image of a ZnSnO3 film viewed along LiNbO3[11¯00]. For this experiment, a stoichiometric LiNbO3 (−Z surface) was used as the substrate; high-quality ZnSnO3 similar to those on congruent LiNbO3 was obtained with identical growth conditions (Fig. S6). It is clear that the film is composed of highly crystallized and c-axis oriented domains over a large area. HAADF and annular bright field (ABF) images around the film/substrate interface are displayed in Figs. 4(b) and 4(c), respectively. In Fig. 4(d), we compared representative patterns, area A in the film and area B in the substrate, together with structural models, indicating the same orientation as observed by XRD. In addition, in the ABF STEM images, where light elements are more clearly resolved than in HAADF STEM ones, it is seen that the asymmetrical contrast due to oxygen ions in the LiNbO3 substrate (area B) is partly maintained in the ZnSnO3 film region (area A). Closely inspecting the image contrast inFigs. 4(a)–4(c), we also find an inhomogeneously bright lattice (area C), which is likely related to crystallographically different nanoclusters. One possible orientation that could explain the irregular pattern is SnO2[001](010)∥LiNbO3[11¯00](0001¯), though this was not detected in the macroscopic XRD characterization presented above. Suppression of such nanoclusters, which might be segregated in the boundaries of two domains of ZnSnO3, would be important for the single-domain formation.

FIG. 4.

(a) HAADF STEM image of a ZnSnO3 film grown on stoichiometric LiNbO3(0001) (−Z surface) taken along LiNbO3[11¯00]. (b) HAADF and (c) ABF STEM images around the interface. (d) Magnified images for area A in the film and area B in the substrate are compared. Schematic structural models along [11¯00] drawn with VESTA24 are shown for comparison. Area C indicated by dashed lines implies local compositional and structural modifications.

FIG. 4.

(a) HAADF STEM image of a ZnSnO3 film grown on stoichiometric LiNbO3(0001) (−Z surface) taken along LiNbO3[11¯00]. (b) HAADF and (c) ABF STEM images around the interface. (d) Magnified images for area A in the film and area B in the substrate are compared. Schematic structural models along [11¯00] drawn with VESTA24 are shown for comparison. Area C indicated by dashed lines implies local compositional and structural modifications.

Close modal

According to first-principles calculation,19,22,23 ZnSnO3 is a wide bandgap semiconductor with a highly dispersed conduction band of Sn 5s character. We determine the optical bandgap Eg of ZnSnO3 and MgSnO3 by transmittance measurement. We adopted LiTaO3(0001) substrates (+Z surface) with a large Eg of 4.6 eV for this measurement because the bandgap of LiNbO3 (3.7 eV) is not sufficiently large as compared to that of the stannate films (Fig. S7). The XRD pattern of the films on LiTaO3(0001) is comparable to that on LiNbO3 (Fig. S3). An extrapolation of the absorption versus photon energy () plot, shown in Fig. 5(a), yields Eg ∼ 3.7 eV for ZnSnO3 and ∼ 4.0 eV for MgSnO3. Although the type of optical absorption is necessary for the detailed analysis, these results underpin the conjectured wide bandgap nature. We then performed photoconductivity measurement for a ZnSnO3 film on LiNbO3(0001) (−Z surface) using Xe lamp irradiation. As shown in Fig. 5(b), the pristine ZnSnO3 sample under dark condition is highly resistive, with a sheet resistance Rs over 1 GΩ at temperature T = 300 K. Photoexcitation at T = 20 K induces persistent photoconductivity [inset ofFig. 5(b); also see Fig. S8] associated with a substantial (several orders of magnitude) decrease in Rs. Hall effect measurement revealed the electron-type carrier, carrier density of 1.5 × 1018 cm−3, and Hall mobility μH of 4.9 cm2 V−1 s−1 at T = 20 K. On heating above T ∼ 200 K, Rs gradually recovers its original high value, indicating the primally electronic origin of the photo-induced conduction.

FIG. 5.

(a) Optical absorbance spectra for ZnSnO3 and MgSnO3 films on LiTaO3(0001) (+Z surface). Absorbance was deduced directly from the measured transmittance without the reflectance correction. For comparison, absorbance values normalized by the film thicknesses were used. (b) Persistent photoconductivity induced by Xe lamp irradiation. The cooling curve in the dark condition was first measured (black curve). At T = 20 K, the temporal evolution of Rs under light illumination (output power P = 240 W) was monitored as shown in the inset (see Fig. S8 for the P dependence). After persistent photoconductivity reached a stable state, the Hall effect was measured. Subsequently, Rs versus T curves are recorded from T = 3 K to 400 K (red curve). Heating above T ∼ 200 K deactivates photoconductivity. (c) T dependence of Rs for a 107-nm-thick ZnSnO3 EDLT at various VG values. The inset shows the T dependence of μH evaluated by the Hall effect measurement.

FIG. 5.

(a) Optical absorbance spectra for ZnSnO3 and MgSnO3 films on LiTaO3(0001) (+Z surface). Absorbance was deduced directly from the measured transmittance without the reflectance correction. For comparison, absorbance values normalized by the film thicknesses were used. (b) Persistent photoconductivity induced by Xe lamp irradiation. The cooling curve in the dark condition was first measured (black curve). At T = 20 K, the temporal evolution of Rs under light illumination (output power P = 240 W) was monitored as shown in the inset (see Fig. S8 for the P dependence). After persistent photoconductivity reached a stable state, the Hall effect was measured. Subsequently, Rs versus T curves are recorded from T = 3 K to 400 K (red curve). Heating above T ∼ 200 K deactivates photoconductivity. (c) T dependence of Rs for a 107-nm-thick ZnSnO3 EDLT at various VG values. The inset shows the T dependence of μH evaluated by the Hall effect measurement.

Close modal

In the photo-excited state, Rs exponentially increases below 20 K. This is possibly due to a carrier localization and a subsequent reduction in μH in the region next to the substrate. Since we previously observed a high field-effect mobility of 45 cm2 V−1 s−1 at T = 300 K in a ZnSnO3 field-effect transistor,17 electrical transport at the film surface under electrostatic charge accumulation is worthy of investigation. To induce a high carrier density, we fabricated an electric-double-layer transistor (EDLT) on the ZnSnO3 surface with an ionic liquid, N,N-diethyl-N-methyl-N-(2-methoxyethyl)ammonium bis(trifluoromethanesulfonyl)imide (See Methods in the supplementary material).Figure 5(c) displays T dependent Rs characteristics at various gate voltages (VG). The application of a positive VG decreases Rs (Fig. S9), as being consistent with the n-type conduction observed in the photoconductivity measurement. A large sheet carrier density close to 1014 cm−2 at VG = 6 V (Fig. S10) induces virtually T-independent behavior with a μH of as high as 20 cm2 V−1 s−1 [inset of Fig. 5(c)], demonstrating the high-mobility nature of the Sn 5s derived conduction band. By fabricating a single-domain ZnSnO3 film, a two-dimensional metallic state may emerge at the polar surface.

In summary, we demonstrated that the use of LiNbO3-type substrates is the key to stabilizing crystalline ZnSnO3 in the thin-film form and discovered a new structural analog MgSnO3 by molecular-beam epitaxy. The results manifest the potential of these wide bandgap stannates for applications in transparent electronics. One of the obvious targets is to fabricate In-free transparent conducting films with appropriate donor elements. A more nontrivial approach is to induce free carriers at the film surface by utilizing its own ferroelectric polarization. Observation of polarization reversal by an electric field15 should be tackled for such studies. The duality of ferroelectric and high-mobility properties in wide bandgap polar stannates may bring ferroelectric materials research into a new stage.

See supplementary material for compositional analysis of Zn-Sn-O films by X-ray photoelectron spectroscopy (Fig. S1), optimization of the Zn/Sn flux ratio based on XRD results (Fig. S2), out-of-plane XRD patterns of ZnSnO3 and MgSnO3 films on congruent LiTaO3(0001) (Fig. S3), wide scan XRD data on congruent LiNbO3(0001) (Fig. S4), substrate dependent growth of MgSnO3 (Fig. S5), out-of-plane XRD patterns of ZnSnO3 on stoichiometric LiNbO3(0001) (Fig. S6), transmittance spectra (Fig. S7), the dependence of persistent photoconductivity on light source power (Fig. S8), transfer characteristics measured at T = 220 K (Fig. S9), and μH data (Fig. S10) for a ZnSnO3 EDLT, for experimental procedures of compositional analysis and device fabrication (Methods in the supplementary material).

The authors thank Y. Inaguma for his insightful advice, T. Matsuoka and S. Kuboya for the use of a spectrophotometer, NEOARK Corporation for the use of a maskless lithography system PALET, and I. Narita, F. Sakamoto, S. Ito, and K. Miura for their experimental assistance. This work was performed under the Inter-University Cooperative Research Program of the Institute for Materials Research, Tohoku University (Proposal Nos. 17G0415 and 17G0417), and also conducted in part at Advanced Characterization Nanotechnology Platform of the University of Tokyo, supported by the “Nanotechnology Platform” of the Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan. This work was supported by JSPS KAKENHI (No. JP15H02022) and Kato Foundation for Promotion of Science (No. KJ-2607).

1.
P.
Shiv Halasyamani
and
K. R.
Poeppelmeier
,
Chem. Mater.
10
,
2753
(
1998
).
2.
K.
Uchino
,
Ferroelectric Devices
, 2nd ed. (
CRC Press
,
Boca Raton
,
2009
).
3.
D. M.
Giaquinta
and
H.-C.
zur Loye
,
Chem. Mater.
6
,
365
(
1994
).
4.
A.
Navrotsky
,
Chem. Mater.
10
,
2787
(
1998
).
5.
Y.
Inaguma
,
M.
Yoshida
, and
T.
Katsumata
,
J. Am. Chem. Soc.
130
,
6704
(
2008
).
6.
A.
Aimi
,
T.
Katsumata
,
D.
Mori
,
D.
Fu
,
M.
Itoh
,
T.
Kyômen
,
K.-I.
Hiraki
,
T.
Takahashi
, and
Y.
Inaguma
,
Inorg. Chem.
50
,
6392
(
2011
).
7.
Y.
Inaguma
,
A.
Aimi
,
Y.
Shirako
,
D.
Sakurai
,
D.
Mori
,
H.
Kojitani
,
M.
Akaogi
, and
M.
Nakayama
,
J. Am. Chem. Soc.
136
,
2748
(
2014
).
8.
D.
Mori
,
K.
Tanaka
,
H.
Saitoh
,
T.
Kikegawa
, and
Y.
Inaguma
,
Inorg. Chem.
54
,
11405
(
2015
).
9.
R.
Yu
,
H.
Hojo
,
T.
Mizoguchi
, and
M.
Azuma
,
J. Appl. Phys.
118
,
094103
(
2015
).
10.
T.
Varga
,
A.
Kumar
,
E.
Vlahos
,
S.
Denev
,
M.
Park
,
S.
Hong
,
T.
Sanehira
,
Y.
Wang
,
C. J.
Fennie
,
S. K.
Streiffer
,
X.
Ke
,
P.
Schiffer
,
V.
Gopalan
, and
J. F.
Mitchell
,
Phys. Rev. Lett.
103
,
047601
(
2009
).
11.
K.
Fujita
,
T.
Kawamoto
,
I.
Yamada
,
O.
Hernandez
,
N.
Hayashi
,
H.
Akamatsu
,
W.
Lafargue-Dit-Hauret
,
X.
Rocquefelte
,
M.
Fukuzumi
,
P.
Manuel
,
A. J.
Studer
,
C. S.
Knee
, and
K.
Tanaka
,
Chem. Mater.
28
,
6644
(
2016
).
12.
M.
Ye
and
D.
Vanderbilt
,
Phys. Rev. B
93
,
134303
(
2016
).
13.
T.
Varga
,
T. C.
Droubay
,
L.
Kovarik
,
M. I.
Nandasiri
,
V.
Shutthanandan
,
D.
Hu
,
B.
Kim
,
S.
Jeon
,
S.
Hong
,
Y.
Li
, and
S. A.
Chambers
,
ACS Appl. Mater. Interfaces
9
,
21879
(
2017
).
14.
T.
Varga
,
T. C.
Droubay
,
M. E.
Bowden
,
L.
Kovarik
,
D.
Hu
, and
S. A.
Chambers
,
Adv. Condens. Matter Phys.
2015
,
493045
.
15.
J. Y.
Son
,
G.
Lee
,
M.-H.
Jo
,
H.
Kim
,
H. M.
Jang
, and
Y.-H.
Shin
,
J. Am. Chem. Soc.
131
,
8386
(
2009
).
16.
D. G.
Schlom
,
L.-Q.
Chen
,
X.
Pan
,
A.
Schmehl
, and
M. A.
Zurbuchen
,
J. Am. Ceram. Soc.
91
,
2429
(
2008
).
17.
H.
Minato
,
K.
Fujiwara
, and
A.
Tsukazaki
,
AIP Adv.
8
,
055327
(
2018
).
18.
JCPDS PDF No.
00-030-0798.
19.
W.-L.
Zhu
,
X.-Y.
Chen
,
Y.-J.
Zhao
, and
T.-S.
Lai
,
Int. J. Mod. Phys. B
28
,
1450224
(
2014
).
20.
A. M.
Marsh
,
S. D.
Harkness
,
F.
Qian
, and
R. K.
Singh
,
Appl. Phys. Lett.
62
,
952
(
1993
).
21.
I.
Tsukada
and
S.
Higuchi
,
Jpn. J. Appl. Phys., Part 1
43
,
5307
(
2004
).
22.
H.
Wang
,
H.
Huang
, and
B.
Wang
,
Solid State Commun.
149
,
1849
(
2009
).
23.
K. P.
Ong
,
X.
Fan
,
A.
Subedi
,
M. B.
Sullivan
, and
D. J.
Singh
,
APL Mater.
3
,
062505
(
2015
).
24.
K.
Momma
and
F.
Izumi
,
J. Appl. Crystallogr.
44
,
1272
(
2011
).

Supplementary Material