Introduction

Heteroepitaxial growth is the epitaxial growth of semiconductor films on dissimilar substrates (e.g., GaN on Al2O3); it has been widely used to obtain single-crystal films for electronic and photonic applications mostly because of the limited availability of single-crystal semiconductor substrates with identical lattice parameters for homoepitaxial growth1,2. Inevitably, heteroepitaxial layers contain high densities of structural defects; because of the lattice mismatch between films and substrates, most of the defects are threading dislocations (TDs), which cross the epitaxial layers perpendicularly2,3. TD cores consist of point defect complexes (e.g., impurities or vacancies) that have abundant dangling bonds, which form deep acceptors with a space-charge double Schottky barrier around dislocation lines; these acceptor-like negatively charged TDs generate an effective potential field that induces scattering of electrons by Coulombic interactions4,5. This “charged TD scattering” severely reduces the number of free charge carriers and thus diminishes electron mobility (μe) in epitaxial semiconductor films (e.g., GaAs4, InN6, GaN3,7). Therefore, the control of charged TD scattering is a key requirement for achieving heteroepitaxial semiconductor films with an electronic grade if homoepitaxy is not available3,8,9.

Charged TD scattering also significantly affects electronic transport in epitaxial BaSnO3 (BSO) films, which are new transparent and wide band-gap oxide semiconductors for transparent and power electronics9,10,11,12. Although La-doped BSO (LBSO) single crystals have excellent an room temperature (RT) μe of ~320 cm2 V−1 s−1 at a carrier concentration of n = 8.0 × 1019 cm−3 and excellent stability, even at high temperatures, heteroepitaxial LBSO films have much lower μe values (at best, ~70 cm2 V−1 s−1 by pulsed laser deposition (PLD)11 and ~124 cm2 V−1 s−1 by molecular beam epitaxy (MBE)13) than single crystals11. This low μe has been attributed to the abundance of TDs9 that inevitably form during epitaxial growth due to large lattice mismatch with the substrate, e.g., SrTiO3 or MgO. Therefore, considerable effort has been devoted to decreasing the density of TDs by using substrates with a lattice constant that is similar to or even the same as that of BSO13,14,15,16 or by inserting buffer layers between BSO epilayers and typical substrates12,17,18,19 to release the lattice mismatch strain.

In this article, we report a new strategy to increase the μe of LBSO epitaxial films by exploiting a recovery process that is assisted by oxygen vacancies. By delicately adjusting both the oxygen partial pressure p(O2) and the temperature during thermal treatment of LBSO epitaxial films, the density of TDs and the degree of mosaicity is significantly reduced, so a maximum RT μe value of 122 cm2 V−1 s−1 at a carrier concentration of n = 1.1 × 1020 cm−3 was achieved. This subtle introduction of an oxygen deficiency during post-treatment improves the crystal quality by decreasing the TD density, NTD, which leads to an increase in mobility with a minimal increase in the number of oxygen vacancies as impurity scattering centers. This new strategy that exploits the interactions between point and line defects can be used to further reduce NTD in heteroepitaxial thin films by using oxygen vacancies to lubricate TDs.

Materials and methods

Synthesis of epitaxial LBSO thin films

Epitaxial La-doped BaSnO3 (LBSO) thin films with a La-doping content of 0–5% and a thickness of ~100 nm were grown on (001)-oriented SrTiO3(001) and MgO single-crystal substrates by using PLD with a KrF excimer laser in a chamber with an \(\sim 10^{ - 6}\) Torr base pressure. The La-doping ratio was controlled by using targets with different La contents; the targets were synthesized by a conventional solid-solution reaction. A KrF excimer laser (λ = 248 nm, τ = 20 ns, Compex Pro 102F, Coherent Inc., Santa Clara, CA) was focused on rotating targets at a fluence of ~1.1 J/cm2 and a repetition rate of 10 Hz. The as-grown LBSO layers were grown in a p(O2) of 260 mTorr at 750 °C to optimize the structural and electrical properties of the as-grown LBSO films.

p(O2) control of LBSO thin films during annealing

The degree of oxygen deficiency in the LBSO films was controlled by adjusting either the p(O2) or the temperature post treatment. The p(O2) post treatment was controlled by simultaneously applying pure Ar gas and 5% forming gas (Ar containing 5% H2) and changing the flow ratio while maintaining a total flow rate of 500 sccm (Fig. S3). To further suppress the reducing ability of H2 gas to achieve subtle control of the p(O2), wet H2 was used by passing H2 gas through a Drechsel bottle that contained water. The H2O ratio in the flowing gas was estimated to be ~3% from the water vapor pressure at RT, i.e., 3.17 kPa at 25 °C. A Yttria-stabilized Zirconia (YSZ) sensor accurately measured log p(O2) as −3.91 in a flow of pure Ar gas, −20.0 in 5% H2 mixed with Ar at 650 °C, and −16.6 in 0.5% wet H2 mixed with Ar at 950 °C. The YSZ sensor measured the voltage difference between the inside of a YSZ cell that was exposed to air and the outside of the same cell that was exposed to flow gas. The measured voltage difference was converted to the p(O2) in the flow gas by using the Nernst equation20:

$$E = \left( {\frac{{{\mathrm{RT}}}}{{4F}}} \right)\log \left( {\frac{{p{\rm{O}}_2\left( {\rm{air}} \right)}}{{p{\rm{O}}_2\left( {{{\rm{flow}}\,{\rm{gas}}}} \right)}}} \right),$$

where E (mV) is the electromotive force, R = 8.3145 (J mol−1 K−1) is the universal gas constant, T (K) is the YSZ cell temperature, F = 96485 (C/mol) is the Faraday constant, pO2(air) = 0.21 (bar) is the oxygen partial pressure of air, and pO2(flow gas) is the oxygen partial pressure of flow gas.

Materials characterization

The crystallinity of the films was characterized using high-resolution X-ray diffraction (HRXRD, Bruker D8 Discovery X-ray diffractometer) with Cu Kα1 radiation (λ = 1.5406 Å) for θ–2θ scan, rocking curve (RC) scan, and reciprocal space mapping (RSM). The surface morphology of the films was observed using an atomic force microscope (AFM, VEECO Dimension 3100). Etching-pit analysis was performed by etching the thin films in 0.4 wt% nitric acid for 10 min before their surface morphologies were observed using AFM. Oxygen 1s X-ray photoemission spectroscopy (XPS) was performed with a synchrotron source in an ultrahigh vacuum chamber (base pressure of the analysis chamber ~5.3 × 10−10 Torr) at the 4D PES beamline at PLS-II of the Pohang Accelerator Laboratory, Pohang, South Korea, to quantify the oxygen deficiencies before and after H2 treatment. The ratio of oxygen vacancies to lattice oxygens was estimated by the areal ratio of the O(II) peak (~531.6 eV) to the O(I) peak (530.1 eV). As a binding energy reference, the Au 4f7/2 peak was set at 84.0 eV. Time-of-flight secondary ion mass spectrometry (TOF-SIMS) depth profiles were recorded by negative polarity with Cs+ sputtering at 3 keV–20 nA.

The resistivity ρ and carrier concentration n were obtained as functions of the temperature from RT (~300 K) to liquid nitrogen temperature (~80 K) using a Hall measurement system with a 0.56-T perpendicular magnetic field. Measurements were made in the van der Pauw geometry with square samples (5 × 5 mm2) and indium ohmic contacts (<1 × 1 mm2) in the sample corners. The four-terminal resistances were measured by a 1-mA current. For the Hall measurements, the magnetic field was varied between ±0.56 T. The carrier concentration was calculated as \(n = 1/\left( {e \times R_{\mathrm{H}}} \right)\), where RH is the measured Hall coefficient. The electron mobility was calculated as \(\mu _e = R_{\mathrm{H}}/\rho\), where the resistivity (ρ = \(R_{\mathrm{S}} \times t\)) was determined by the van der Pauw method and is independent of any errors in determining the film thickness.

Theoretical estimation of the dislocation-limited mobility

The dislocation-limited mobility, \(\mu _{e,{\mathrm{TD}}}\), was estimated by varying the TD density in the equation3:

$$\mu _{\mathrm{{e}},{\mathrm{TD}}} = \frac{{30\sqrt {2\pi } \varepsilon _0^2\varepsilon _{\mathrm{s}}^2a^2\left( {kT} \right)^{\frac{3}{2}}}}{{N_{{\mathrm{TD}}}e^3f^2{\mathrm{\lambda }}_{\mathrm{D}}\sqrt {m_{\mathrm{e}}^ \ast } }},$$

where

$${\mathrm{\lambda }}_{\mathrm{D}} = \left( {\frac{{\varepsilon _0\varepsilon _skT}}{{ne^2}}} \right)^2.$$

In these equations, \(\varepsilon _0\) = 8.854 × 10−12 (F/m) is the vacuum dielectric constant, \(\varepsilon _{\mathrm{s}}\) = 21 is the relative dielectric constant12, a =  4.116 (nm) is the lattice constant, k = 1.381 × 10−23 (J/K) is the Boltzmann constant, T = 300 (K) is the temperature, NTD (cm−2) is the TD density, e = −1.602 × 10−19 (C) is the electron charge, f = 1 is the fraction of occupied acceptor centers of each TD, \(m_{\mathrm{e}}^ \ast\) = 0.1731 × 10−31 (kg) is the electron effective mass12, and n (cm−3) is the carrier concentration.

Results and discussion

Epitaxial LBSO ([La] = 0–5%) thin films with a thickness of 100 nm were grown on (001)-oriented SrTiO3 (STO) and MgO substrates by PLD at 750 °C. HRXRD θ–2θ scans of the as-grown films showed only the (00l) reflection; this result indicates that the films had (001) pseudomorphic epitaxy without secondary phases. However, the lattice mismatch of BSO (a = 0.4116 nm) is −5.4% with the (001)-oriented STO substrate (a = 0.3905 nm) and +2.2% with the (001)-oriented MgO (a = 0.4212 nm) substrate; these mismatches cause misfits and TDs to form in the epitaxial LBSO layers. The as-grown 1% LBSO epitaxial films exhibited the parameters n ~5.3 × 1019 cm−3 and RT μe ~15 cm2 V−1 s−1 when grown on STO and n ~4.1 × 1019 cm−3 and RT μe ~16 cm2 V−1 s−1 when grown on MgO (Fig. 1, Fig. S1); these values are comparable to those of other as-grown 1% LBSO epitaxial films produced by PLD (n ~6 × 1019 cm−3, μe ~20 cm2 V−1 s−1)11,21. When 1% La dopants are fully activated, n is expected to be ~1.5 × 1020 cm−3, so the observed value of n ~6 × 1019 cm−3 indicates that approximately only one-third of the La3+ dopants provide conducting electrons. This low yield suggests that the donated electrons are trapped at the TDs.

Fig. 1: Increase in electron mobility of LBSO epitaxial thin films after precise H2 treatment.
figure 1

Temperature-dependent (a) electrical resistivity ρ, (b) carrier concentration n, (c) and electron mobility μe of 1% La-doped BaSnO3 thin films: as-grown (black-filled squares), Ar-annealed at 950 °C (green-hollow squares), 5% H2-annealed at 650 °C (red-filled circles), and wet 0.5% H2-annealed at 950 °C (blue-hollow circles). (d) μe as a function of n at room temperature (RT) in LBSO films with different La-doping contents: as-grown (black-filled squares), 5% H2-annealed at 650 °C (red-filled circles), and wet 0.5% H2-annealed at 950 °C (blue-hollow circles). Dashed lines: theoretically calculated μe limited by charged-dislocation scattering as a function of threading dislocation density, NTD. Note that the RT μe maximum of the H2-treated LBSO films shifts toward the upper left corner and is located near 1020 cm−3 (orange arrow)

The \(\mu _e\) of the as-grown epitaxial layer was optimized using various post-treatments to tune the oxygen chemical potential of the samples. As an initial attempt, the samples were annealed under inert Ar at 650 or 950 °C. Annealing at 650 °C did not change n or μe, but annealing at 950 °C increased μe to ~30 cm2 V−1 s−1 at RT with an almost constant n; this result is consistent with previous reports that μe can be increased by annealing in inert N2 at 1000 °C21.

To further modulate the degree of oxygen deficiency during the thermal treatment of as-grown LBSO epitaxial films with large TD densities, the LBSO films were thermally annealed at 650 °C under a reducing atmosphere with 5% H2-mixed forming gas (denoted 5% H2); both n and the RT μe increased (μe ~90 cm2 V−1 s−1 at n ~1.2 × 1020 cm−3). The increase in n under a reducing atmosphere may be related to the introduction of carrier-generating point defects such as oxygen vacancies or interstitial hydrogen22,23,24. Treatment under excessively reducing conditions at high temperature (i.e., 5% H2 at 950 °C) to increase the diffusivity of the point defects led to a significant suppression of the (001) XRD peak and insulating properties of the LBSO epitaxial films; these changes can be attributed to the collapse of the BSO crystal structure due to excessive formation (or incorporation) of oxygen vacancies in BSO (Fig. S2). Therefore, to exploit the high ionic mobility at high temperature without causing the collapse of the BSO crystal structure, the concentration of generated oxygen vacancies should be regulated to maintain the lattice framework during H2 treatment.

To achieve this balance, the p(O2) was precisely controlled in the following way. By flowing 0.5% H2 gas through a Drechsel bottle that contained water, H2O gas was added to the H2 gas in the furnace (i.e., wet H2) (Fig. S3)25. The effects of this experimental setup could be predicted from the following reaction equation:

$${{{\rm{BaSnO}}_3}}\left( {\mathrm{s}} \right) + {{{\rm{H}}_2}}\left( {\mathrm{g}} \right) \rightleftharpoons {\rm{BaSnO}}_{3 - \delta }\left( {\mathrm{s}} \right) + {{{\rm{H}}_2O}}\left({\mathrm{g}}\right).$$

Based on Le Chatelier’s principle, both the introduction of H2O and a decrease in the H2 content would suppress the formation of H2O and oxygen vacancies in BaSnO3, even at high temperature. Measurements using an oxygen sensor based on YSZ detected a higher p(O2) in wet H2 at 950 °C (log p(O2) = −16.6) than in 5% H2 at 650 °C (log p(O2) = −20.0)20. This difference confirms that the utilization of wet H2 is an effective method for slightly suppressing the reducing ability at high temperature; this suppression cannot be achieved using 5% H2 alone. With the accurate tuning of the oxygen chemical potential, the wet H2 process at 950 °C gave the maximum RT μe of LBSO (~122 cm2 V−1 s−1 at n ~1.1 × 1020 cm−3) without collapsing the lattice framework of LBSO, even at high temperature. Therefore, the optimized thermal treatment with subtle p(O2) control could remarkably increase the RT μe in LBSO thin films with a minimal increase in the number of oxygen vacancies as impurity scattering centers. The thermal H2 treatment also increased μe and n in the LBSO epitaxial films on MgO substrates (Fig. S1: μe and n increased from 16 cm2 V−1 s−1 and 4.1 × 1019, respectively, in as-grown LBSO to 76 cm2 V−1 s−1 and 7.3 × 1019 in wet H2 LBSO at 950 °C (50 cm2 V−1 s−1 and 8.1 × 1019 in 5% H2 LBSO at 650 °C) despite the severe crystal mosaicity in LBSO on MgO (Fig. S4); this result indicates that the wet H2 process is very effective for increasing the RT μe regardless of the substrate. To confirm the chemical origin of the point defects and generate more carriers after the H2 treatments, we performed TOF-SIMS and XPS on the as-grown and H2-treated BSO epitaxial films (Fig. S5). The hydrogen anion (H) signal increased very little after the H2 treatments (Fig. S5a)26,27, so hydrogen interstitials are unlikely to be the main origin of mobile carriers. Instead, deconvolution of the O 1s XPS spectra of both films indicates that the areal ratio of the O(II) peak (~531.6 eV) to the O(I) peak (~530.1 eV), which are related to oxygen vacancies and lattice oxygens23, respectively, increased from 0.588 in as-grown BSO to 1.42 in the 5% H2 BSO films (Fig. S5b, c); this result reveals the formation of oxygen vacancies, which can be additional sources of mobile electrons in the LBSO films after H2 treatment. However, oxygen vacancies usually scatter carriers and thus cannot be the origin of the RT μe increase. Indeed, the formation of oxygen vacancies decreases the RT μe of the LBSO thin film due to increased scattering by the charged point defects28, which is contrary to our present result in which RT μe increased even with the formation of oxygen vacancies after H2 treatment; this contradiction suggests that a different origin induces the increase in μe after H2 treatment.

To exclude the possibility of electrical conduction through the STO substrate after our precise H2 treatment, the substrate conductance was indirectly estimated from the linear extrapolation of the sheet conductance and (Fig. S6) directly measured with respect to the H2 treatment at different p(O2) values and temperatures (Fig. S7). Unlike in the typical reducing conditions for producing oxygen-deficient SrTiO3−δ, the STO substrates showed negligible conduction comparable to that of the pristine STO substrate after our delicate thermal treatment (i.e., 5% H2 LBSO at 650 °C and wet H2 LBSO at 950 °C). Therefore, it can be inferred that our mild reducing conditions produced oxygen vacancies in only the LBSO epitaxial layers and are not sufficient to produce oxygen deficiency in the STO substrates. This selective formation of oxygen vacancies in LBSO/STO can be explained by the fact that the formation energy of oxygen vacancies in STO (~3.1 eV)29 is considerably higher than that in BSO (~1.7 eV)24.

To gain further insight into electron transport in the LBSO epitaxial films, we measured the temperature dependence of n and μe for 1% La-doped LBSO films treated under different conditions (Fig. 1b, c). n was almost constant with varying temperature in all samples (Fig. 1b), indicating that the films are degenerately doped. In contrast, the temperature dependence of μe was affected by the p(O2) used during post-treatment (Fig. 1c); this relationship indicates that the dominant carrier scattering mechanism changed. Samples without H2 treatment showed no temperature dependence of μe (e.g., 15 cm2 V−1 s−1 at 300 K and 17 cm2 V−1 s−1 at 80 K for as-grown LBSO), but those subject to H2 treatment showed a significant increase in μe as the temperature decreased (e.g., 122 cm2 V−1 s−1 at 300 K and 197 cm2 V−1 s−1 at 80 K for the LBSO film annealed in wet H2 at 950 °C).

To understand the significant increase in μe after the H2 treatment, the resulting μe can be expressed according to Matthiessen’s rule3,12,17:

$$\mu _{\rm e}(T)^{ - 1} = \mu _{{\rm e},{\rm{LO}}}(T)^{ - 1} + \mu _{{\rm e},{\rm{CI}}}(T)^{ - 1} + \mu _{{\rm e},{\rm{TD}}}(T)^{ - 1},$$

where μe,LO is μe governed by LO phonon scattering, μe,CI is μe governed by charged impurities, and μe,TD is μe governed by TDs. The dominant scattering factor changes with NTD, as previously reported in the plot of the temperature dependence of \(\mu _{\mathrm{e}}(T)\)12,17: when NTD is high, the contribution of weakly temperature-dependent μe,TD is dominant over the contribution of μe,LO and μe,CI, so the resultant μe exhibits negligible change with temperature. However, if NTD is low, μe,TD is negligible compared to μe,LO and μe,CI, so the resultant μe shows a strong temperature dependence due to LO phonon-limited scattering, charged-impurity-limited scattering, or both. Our analysis of the dominant scattering mechanisms suggests that the precise tuning of oxygen vacancies in our LBSO films induces a change in the scattering processes from charged-dislocation-limited scattering to ionized impurity- or LO phonon-limited scattering (or both) by reducing NTD.

The change in the dominant scattering mechanism upon H2 treatment can be readily seen in the plot of the RT μe of the LBSO films ([La] = 0–5%) as a function of n at RT (Fig. 1d). Due to competition between charged-dislocation scattering and charged-impurity scattering, the plot of the RT μe vs. n is bell-shaped3,9,12,17. In the as-grown and Ar-treated LBSO films, μe was maximized at high n (~5 × 1020 cm−3) and then decreased with decreasing n; this relationship demonstrates the strong influence of charged dislocations as charge traps, scattering centers for carriers, or both, and also shows that this influence reduces n and μe simultaneously. After H2 treatment of the LBSO films, both n and μe increased simultaneously in all the LBSO films; this change led to a shift in each data point of n vs. μe toward the upper right corner. This shift indicates that the H2 treatment causes a decrease in the intensity of charge TD scattering despite the increase in the number of charged point defects. In the overall bell-shaped curves, the maximum μe of the H2-treated LBSO films shifted toward the upper left corner (Fig. 1d, orange arrow) and is located near 1020 cm−3. In GaN epitaxial films with a high NTD, the shift in the maximum μe toward the upper left corner indirectly represents a decrease in NTD3, which again confirms that our subtle H2 treatment reduces NTD and the contribution of charged-dislocation scattering.

As a further check of our model, various X-ray diffraction techniques, e.g., RSM around the (103) Bragg reflection and RC of the (002) reflection, were used to explore the structural origin of the RT μe increase induced by the p(O2)-controlled thermal treatment (Fig. 2). At first glance, the (103) Bragg reflections of the precisely H2-treated LBSO films (Fig. 2c, d) had higher peak intensities and narrower peak widths than those of the as-grown (Fig. 2a) LBSO films (Fig. 2b). Ar treatment of the LBSO films had almost no effect on the full-width half maximum (FWHM) in both the RSM and RC, but both H2 treatments caused significant reductions in the FWHMs in the in-plane direction (ΔQx) and the out-of-plane direction (ΔQz). In particular, ΔQx decreased considerably more than ΔQz and the RC Δω in the H2-treated LBSO films (Fig. 2e) (e.g., ΔQx = 0.0172 Å−1\(\to\)0.0111 Å−1, ΔQz = 0.0099 Å−1\(\to\)0.0088 Å−1, Δω = 0.033°\(\to\)0.026° in the wet 0.5% at 950 °C H2 samples). In contrast to the increase in the number of oxygen vacancies in the BSO films after H2 treatment, the distinct narrowing of the RSM (i.e., ΔQx) peaks after thermal treatment under a controlled p(O2) should be related to structural improvement by the reduction in the density of line defects (e.g., \(N_{{\rm{TD}}}\)) in the films, which is likely to be the main origin of the increase in the RT μe.

Fig. 2: Structural improvement of LBSO epitaxial films after H2 treatment.
figure 2

Reciprocal space mapping (RSM) around (103) reflections from LBSO epitaxial films: (a) as-grown, (b) Ar-annealed at 950 °C, (c) 5% H2-annealed at 650 °C and (d) wet 0.5% H2-annealed at 950 °C. e Full-width at half-maximum of the (103) RSM peak along the x direction (ΔQx) and z direction (ΔQz) and of the (002) rocking curve (Δω) for the LBSO films in ad. The RSM peak width narrowed substantially along the Qx direction after H2 treatment of the LBSO films

A reduction in NTD for a specific type of TD can explain the reduction in the RSM FWHM in ΔQx after H2 annealing. Dislocations cause local lattice displacement, u(r), which influences the X-ray scattering intensity I(q) as follows:

$$I\left( q \right) \propto \left| {{\int} {e^{ - iq \times u\left( r \right)}e^{ - iq \times r}dr} } \right|^2,$$

where q is a diffraction vector that corresponds to the Bragg reflection plane (hkl)30. Therefore, if lattice distortion is induced in the in-plane direction (u1 = (x, 0, 0)), then I(q + \(\Delta\)qx) increases but I(q+\(\Delta\)qz) does not, so peak broadening occurs along the x direction. In the case of TDs, in-plane distortion, u1, can be induced by edge-type dislocations but not by screw-type dislocations30,31. Therefore, the type of TD in the epitaxial layers affects the direction of lattice distortion, so the RSM FWHMs change asymmetrically. For example, in GaN epitaxial films with a high density of TDs, edge-type TDs widen the in-plane FWHMs and screw-type TDs widen the out-of-plane FWHMs30,31,32,33.

Likewise, the observed reduction in the in-plane FWHMs in our LBSO epitaxial films should be related to a decrease in the edge-type NTD rather than the screw-type NTD after the H2 treatment. Therefore, the H2 atmosphere at high temperature accelerated the elimination of edge-type TDs, which are expected to be the dominant TD type in the as-grown LBSO epitaxial films. Indeed, a Burgers vector of <010>a was mostly observed in the as-grown LBSO thin films by two-beam dark-field transmission electron microscopy (TEM) imaging (Fig. S8); this observation confirms that edge-type TDs were the dominant TD type. The reduction in NTD after the H2 treatment was confirmed by etching-pit analysis using AFM (Fig. S9). Etching with 0.4% nitric acid creates pits near the cores of the TDs because etching is faster near defective and deformed atomic structures than nearly ordered ones21,34. The H2-treated film with a narrow ΔQx had far fewer etching pits than the as-grown film, i.e., NTD was lower in the H2-treated epilayers than in the as-grown epilayers.

The importance of the precise p(O2) control for inducing the structural improvement responsible for the increases in the RT μe can be demonstrated by the changes in the structural and electrical properties of undoped BSO films after systematically controlling the p(O2) to change the percentage of H2 gas (0–10 %) at 650 °C during thermal treatment for 1 h (Fig. 3). As the amount of H2 gas increased (p(O2) decreased) during post-annealing, ΔQx and the RT resistivity of the annealed BSO drastically decreased and almost saturated when the H2 content was 1%; this trend indicates that the improvement in the in-plane structural quality (ΔQx), not ΔQz, seems to be strongly correlated with the increase in electrical transport. The best structural quality can be observed in the range 0.4 ≤ H2 content < 1% H2, but an H2 gas content >5% H2 caused ΔQx to increase again; this result confirms that the excessive generation of oxygen vacancies degrades the overall crystal quality. Therefore, simultaneous increases in both the structural and electrical properties require that the p(O2) be precisely controlled. This conclusion may explain why the RT μe can be increased using the wet H2 process to slightly improve the precision of the p(O2) control. If this is the case, the increase in μe after H2 treatment may occur because the improvement imparted by the contribution of μTD dominates the degradation caused by the contribution of μCI.

Fig. 3: The importance of the precise control of the p(O2) for increasing the RT μe through structural improvement.
figure 3

a Electrical resistivity (black square), carrier concentration (red circle), electron mobility (blue triangle) and (b) corresponding RSM FWHM ΔQx and ΔQz of BSO thin films H2-annealed at 650 °C with changing H2/Ar ratio. Note that careful control of the p(O2) is essential for achieving simultaneous increases in the structural and electrical properties, which can explain the further improvement in the RT \(\mu _{\rm{e}}\) at the small expense of point defect generation by precise p(O2) control

The electrical and structural results both suggest that the formation of oxygen vacancies by the precise H2 treatment reduces the density of TDs and thereby increases the RT μe. Dislocation annihilation assisted by point defects has been traditionally considered an effective recovery process in physical metallurgy35: This recovery process is exploited in single-crystal-like materials to release strain energy concentrated around dislocations by removing or rearranging structural defects during the annealing of highly deformed samples at high temperature. The recovery process during heat treatment is kinetically limited by atomic migration, so point defects (e.g., oxygen vacancies in the present study) facilitate dislocation motion and efficiently increase the release of deformed strain35,36,37. In our case, the LBSO epitaxial films grown by PLD contain extended defects, e.g., TDs, which are either in equilibrium or non-equilibrium; the films also contain local deformations near the TD cores due to lattice/thermal mismatch, high-energy growth processes, or both.

Along the dislocation lines, the columns of grown films are slightly tilted or mis-oriented relative to each other, forming mosaic-like structures with local deformation9. Once sufficient thermal energy is supplied, the deformation energy can be released in a recovery process that entails dislocation annihilation, thereby reducing NTD and decreasing the lattice distortion induced by the TDs. Because dislocation movement is a prerequisite for dislocation annihilation, point defects such as oxygen vacancies accelerate the movement of dislocations along and across the slip plane, e.g., dislocation climb35,36. This climb occurs by the absorption or emission of vacancies, which are thermally activated processes35, so the rate of climb is directly proportional to the vacancy concentration in ionic crystals. Therefore, dislocation movement that is assisted by oxygen vacancies without creating excessive oxygen deficiencies appears to be very important for improving the structural and electrical properties of the LBSO epitaxial layer by reducing NTD. A similar improvement in the structural quality by vacancy-dislocation interaction has been reported in cold-worked Al, in which the presence of vacancies significantly increased the mobility of dislocations36.

Finally, Fig. 4 shows a comparison of the variation in μe with n at RT for our H2-treated LBSO/STO films with various La concentrations and previously reported LBSO/STO films. With the exception of the bulk LBSO crystals, in which μe increased a relatively small amount as n decreased, all the LBSO films showed bell-shaped curves in the plot of μe vs. n at RT. This curve represents charged-dislocation-limited scattering at low n and ionized impurity scattering at high n, at which charge screening at high n suppresses dislocation scattering as a dominant scattering mechanism. The maximal RT μe as a function of n was significantly shifted toward the upper left corner in our wet H2-treated LBSO films compared to the other as-grown LBSO films prepared by PLD. This shift indicates that our precise H2 treatment significantly reduced NTD. This reduction in NTD induced by the optimized control of the interactions between point defects and line defects achieves a maximal RT μe in our LBSO (~122 cm2 V−1 s−1 at ~n = 1.1 × 1020 cm−3) that is comparable to the RT μe of other LBSO films whose properties were improved by buffer layers, homoepitaxial growth, or different growth methods (e.g., MBE). However, at a high n > 1020 cm−3, μe significantly decreased with increasing n in our case, but the LBSO thin films grown by MBE with controlled cation stoichiometry17 maintained a relatively high RT μe in spite of the increase in n and thus achieved a low resistivity <10−4 Ω cm by minimizing the impurity scattering induced by the cation off-stoichiometry. Therefore, a further improvement in the RT μe of our H2-treated LBSO through the precise control of the residual point defects and line defects as scattering centers is possible.

Fig. 4
figure 4

Comparison of the room temperature electron mobility vs. carrier concentration for our optimized LBSO films wet 0.5% H2-annealed at 950 °C (red stars) and previously reported LBSO films, i.e., PLD-grown LBSO thin films on an STO substrate11 (black squares), and on a single-crystal BSO substrate16 (green circles), MBE-grown LBSO thin films on an STO substrate13 (blue triangles, orange diamonds), and single-crystal LBSO38 (pink hexagons)

Conclusion

The precise control of the oxygen partial pressure p(O2) and temperature during post-treatment with H2 increased n and μe simultaneously in epitaxial La-doped BaSnO3 (LBSO) films and thereby achieved a significant increase in the RT μe (122 cm2 V−1 s−1 at a carrier concentration of 1.1 × 1020 cm−3). Despite the possibility that the H2 treatment could increase the number of scattering centers caused by charged point defects (i.e., oxygen vacancies), the precise control of the p(O2) enabled an increase in the mean free path of electrons in the LBSO thin films. This change is mostly attributed to a reduction in the number of TDs and the mosaicity by the accelerated dislocation movement that occurs due to oxygen vacancy-assisted recovery in ionic crystals. These results suggest that precise control of the interactions between point and line defects can be used to switch the dominant carrier scattering centers in heteroepitaxial oxide films. This strategy using defect chemistry may provide new approaches for achieving perovskite oxide heterostructures that have a high RT μe.