The influence of heat treating n-type bulk β-Ga2O3 in hydrogen (H2) and argon (Ar) gases on the presence of the defect level commonly labeled as E1 was studied. Fourier transform-infrared spectroscopy confirms that hydrogen (H) is incorporated into β-Ga2O3 during H2 annealing at 900 °C. Deep-level transient spectroscopy measurements reveal that the concentration of the E1 level is promoted by the introduction of H, in contrast to what is observed in samples heat-treated in an Ar flow. We further find the E1 level to be stable against heat treatments at 650 K, both with and without an applied reverse-bias voltage. Potential candidates for the defect origin of E1 are investigated using hybrid-functional calculations, and three types of defect complexes involving H are found to exhibit charge-state transition levels compatible with E1, including substitutional H at one of the threefold coordinated O sites, Ga-substitutional shallow donor impurities passivated by H, and certain configurations of singly hydrogenated Ga–O divacancies. Among these types, only the latter exhibit H binding energies that are consistent with the observed thermal stability of E1.

Monoclinic gallium sesquioxide (β-Ga2O3) is an ultra-wide bandgap semiconductor (Eg 4.9 eV 1–3) that has shown promise for applications in power electronics and UV photodetectors.4–9 For β-Ga2O3 to live up to its potential, it is important to control the electrically active defects in the material since defects play a crucial role in determining the electrical conductivity of a semiconductor by acting as dopants or compensating centers.10 Furthermore, defects can influence the operation of β-Ga2O3-based devices by, e.g., pinning the Fermi level11–13 or acting as recombination centers.14 Consequently, defect levels have been studied to a great extent in β-Ga2O3, using techniques such as electron paramagnetic resonance,15–17 cathodoluminescence,18,19 steady-state photo-capacitance,20–30 and deep-level transient spectroscopy (DLTS).11,12,22–24,26,31–36

Recently, H-related defects in β-Ga2O3 have attracted considerable attention.37–41 It has been shown in experimental and computational studies that Ga vacancies (VGa) complexed with H are likely to form in n-type material in the presence of H42,43 and are expected to exhibit deep defect levels.33,43 There are also a number of reports that propose H to be associated with shallow donor states,38–40,44 potentially due to the formation of interstitial H (Hi) or H substituting for O (HO).45 Several other H-related defects have also been reported.36,39,46–48

The E1 center is a DLTS defect signature with an activation energy of about 0.6 eV that has been observed previously in as-received bulk crystals grown by edge-defined film-fed growth (EFG) and the Czochralski (CZ) method,31–33 as well as in epitaxial layers grown by molecular beam epitaxy (MBE)49 and halide vapor phase epitaxy (HVPE).23 Polyakov et al. observed an increase in the concentration of E1 when subjecting EFG-grown bulk crystals with a surface orientation of (010) to a H-plasma,50 whereas Irmscher et al. showed that the concentration of E1 in CZ-grown bulk crystals was not increased by a high-temperature heat treatment in O2 ambient.31 

Different reports exist regarding the effect of irradiation on the E1 level. Ingebrigtsen et al. and Farzana et al. did not observe any change in the E1 concentration in EFG-grown bulk crystals, following 0.6 and 1.9 MeV proton,33 and neutron irradiation,28 suggesting that E1 cannot be solely related to intrinsic defects. In contrast, Polyakov et al. reported a slight increase in the E1 concentration following 20 MeV proton irradiation,24 18 MeV α-particle irradiation,24 and pulsed fast reactor neutron irradiation51 of HVPE films.

Here, we report on the effect of H2 and Ar annealing on the E1 level in EFG-grown bulk β-Ga2O3 crystals. The introduction of H into the crystals by annealing in H2 is confirmed through Fourier transform-infrared spectroscopy (FT-IR) measurements, which reveal an O–H vibrational line previously assigned to a doubly hydrogenated Ga vacancy (VGaib2H).42 From DLTS measurements, we find that H2 heat treatments at 900 °C promote the concentration of E1, whereas equivalent heat treatments performed in an inert Ar flow do not generate any notable changes in the E1 concentration. We further find that the charge-carrier concentration is not influenced by H2 heat treatments. Annealing at 650 K with and without an applied reverse-bias voltage revealed that the E1 center is stable under these conditions. Finally, we discuss potential defect origins of the E1 center based on comparison with hybrid-functional calculations on H-related defects in β-Ga2O3.

Bulk EFG-grown β-Ga2O3 crystals52,53 with a surface orientation of (201) were purchased from Tamura Corporation,54 including two different 0.7 mm thick wafers originating from different production batches. Both wafers were unintentionally doped n-type. The wafers were cut into samples measuring approximately 5×5 mm2 using a laser cutter.

Some of the samples were subjected to heat treatments in closed quartz ampoules filled with approximately 0.5 bar of H2 at room temperature. The ampoules containing the samples were evacuated with a roughing pump prior to filling with H2. Particularly, three cycles of evacuation and filling with H2 were performed before eventually filling the ampoule with 0.5 bar of H2 and subsequently sealing the ampoule. The heat treatments were performed in a tube furnace at a temperature of 900 °C for an annealing duration (tann) of 15–75 min. Once the furnace reached the set temperature, the ampoule containing the sample was put into the tube furnace, annealed for the desired duration, and then removed from the furnace to cool down. The samples were removed from the ampoules after they had reached room temperature; i.e., the samples were not exposed to air at elevated temperatures. In addition, three samples were subjected to heat treatments in an Ar flow at the same temperature of 900 °C for a tann between 15 and 60 min.

Using FT-IR, infrared absorbance spectra were measured at 5 K on as-received and both H2- and Ar-annealed samples. The measurements were performed utilizing a Bruker IFS 125HR spectrometer equipped with a globar light source, a KBr beam splitter, and an InSb detector. The samples were cooled in a Janis PTSHI-950-5 closed-cycle, low vibration pulse tube cryostat filled with He exchange gas and equipped with ZnSe windows. All measurements used a spectral resolution of 0.5 cm−1 with unpolarized light incident along the direction normal to the (201) surface of the crystals. The single-channel spectrum of the empty sample holder was used as a reference. The recorded transmittance (Tr) data were converted to absorbance (A) using the equation A=log10(Tr).55 

For the electrical characterization, Schottky barrier diodes (SBDs) were fabricated on samples in the as-received state or after H2/Ar annealing. Circular Ni pads with diameters between 300 and 900 μm were deposited using e-beam evaporation and a shadow mask.32,33,56 Typically, a contact thickness of 150 nm was used. Stacks of Ti (thickness = 10 nm) and Al (thickness = 150 nm) were used as Ohmic contacts covering the back side of the samples.

Current–voltage (IV) and capacitance–voltage (CV) measurements were performed in the dark on all SBDs to ensure that the devices were suitable for DLTS. IV measurements were performed using a Keithley 6487 picoammeter/voltage source, whereas CV measurements were conducted using a Boonton 7200 capacitance meter or an HP 4280A capacitance meter. From CV measurements, using a probing frequency of 1 MHz, the donor concentration (ND) of the samples was determined10 assuming a static dielectric constant of 10.2.57 Moreover, the widths of the space-charge region (W), and hence the probing depths for DLTS measurements, were estimated from the CV measurements.10 

DLTS measurements were performed on two setups, which both are refined setups of the one described in detail in Ref. 58, covering the temperature range from 150 to 700 K. The DLTS spectra were constructed using a GS2 filter (lock-in filter).59 The spectra are displayed as 2NDΔC/Crb, where ΔC denotes the amplitude of the capacitance transient measured in DLTS, whereas Crb represents the quiescent capacitance of the SBD at the applied reverse-bias voltage.10,60 Parameters describing the electron traps observed in DLTS measurements, such as the trap concentration (Nt), the activation energy (EA), and the apparent capture cross section (σna), were obtained by comparing the recorded DLTS spectra with simulations using a python-based script.20 Here, Nt was computed by taking the λ-correction into account.10,20,31 The uncertainty in EA is estimated to be around 0.04 eV, whereas the uncertainty in σna can be expected to be within ± one order of magnitude.35 

To gauge the stability of the E1 center, heat treatments of the 60 minH2-annealed sample were performed. The annealing was conducted up to 650 K with and without an applied reverse-bias voltage of −5 V, denoted as reverse-bias annealing (RBA) and zero-bias annealing (ZBA), respectively. The annealing cycles were performed in the same manner as described in Ref. 36, except for a slower heating rate of 5 K/min and an HP 4280A capacitance meter as the voltage source.

First-principles calculations were performed using the projector augmented wave method61,62 and the Heyd–Scuseria–Ernzerhof (HSE)63 screened hybrid functional, as implemented in the VASP code.64 The Ga 3d electrons were included in the valence, and the fraction of screened Hartree–Fock exchange was adjusted to 33%. This results in a direct bandgap value of 4.9 eV and lattice parameters (a=12.23Å, b=3.03Å, c=5.79Å, and β=103.8°) in good agreement with experimental data.1,65 Defect calculations were performed using 160-atom supercells, a plane-wave cutoff of 400 eV, and a single special k-point at (0.25, 0.25, 0.25). Defect formation energies and thermodynamic charge-state transition levels were evaluated using the formalism described in Ref. 66, with finite-size corrections applied for charged defects.67–69 Binding energies of H-related defect complexes were calculated as the difference between the formation energy of the complex and the sum of the formation energies of Hi and the remaining entity when one H is removed from the complex.66 A positive binding energy indicates a stable complex.

To facilitate comparison between the hybrid-functional calculations and EA extracted for E1 from DLTS data, we have constructed one-dimensional configuration coordinate (CC) diagrams describing the dynamics of the electron capture and emission process.70–72 CC model parameters were derived from the hybrid-functional calculations, including the ionization energy (Ei), the change in the configuration coordinate (ΔQ), and the ground and excited state Franck–Condon shifts (dgFC and deFC). EA extracted from DLTS includes an energetic barrier for electron capture (Eb) in addition to Ei. In the CC model, this barrier is obtained from the intersection point of the potential energy curves in the ground and excited state.71 

Figure 1 shows baseline-corrected IR absorbance spectra recorded on as-received and H2-annealed samples. The baseline in the absorbance spectra originates from surface reflection losses, scattering at the rough back surface, and free charge-carrier absorption.42 The samples annealed in H2 exhibit an absorbance feature at around 3437 cm−1, which is related to a localized vibrational mode (LVM) associated with VGaib2H.42 Indeed, VGaib2H has previously been found to form under n-type conditions during H2 annealing,42 in line with first-principles calculations.33,43 The as-received and Ar-annealed samples did not show such an absorbance feature (FT-IR data for samples annealed in Ar are not shown), indicating that VGaib2H is only present in negligible amounts in the bulk of these samples. The data were modeled with Lorentzian profiles to compute the integrated absorbance of the feature related to the LVM of VGaib2H. The integrated absorbance is proportional to the concentration of VGaib2H in the bulk crystals, and hence, an approximately linear relation between the VGaib2H concentration and tann can be seen in the inset of Fig. 1. Thus, the results displayed in Fig. 1 show that H penetrates into the bulk of β-Ga2O3 during H2 annealing. However, the concentration of H is too low to be measured by, e.g., chemical techniques, such as secondary ion mass spectrometry. Note that the small shoulder that can be discerned at 3439.5 cm−1 in Fig. 1 is caused by noise in the baseline.

IV and CV measurements on SBDs comprising as-received, H2-annealed, and Ar-annealed samples showed that the SBDs were suitable for performing DLTS measurements. In IV measurements, SBDs fabricated on H2-annealed samples typically displayed a larger leakage current (current under an applied reverse-bias voltage) compared to as-received and Ar-annealed samples. This limited the tann that could be used for the H2 heat treatments. The increase in leakage current might be related to roughening of the sample surface during the H2 annealing.39,50

FIG. 1.

Baseline-corrected IR absorbance spectra recorded on as-received and H2-annealed EFG β-Ga2O3 bulk crystals. The data (full circles) were modeled with Lorentzian profiles (solid lines) to extract the integrated absorbance of the feature at around 3437 cm−1, which is related to a LVM associated with VGaib2H.42 The inset shows the dependence of the integrated baseline-corrected absorbance on tann.

FIG. 1.

Baseline-corrected IR absorbance spectra recorded on as-received and H2-annealed EFG β-Ga2O3 bulk crystals. The data (full circles) were modeled with Lorentzian profiles (solid lines) to extract the integrated absorbance of the feature at around 3437 cm−1, which is related to a LVM associated with VGaib2H.42 The inset shows the dependence of the integrated baseline-corrected absorbance on tann.

Close modal

From CV measurements, ND values between 2 × 1017 and 5×1017 cm−3 were determined for all samples independent of the heat treatment. The values determined for ND indicate that the Fermi level is close to EC in all investigated samples. From CV measurements, the typical probing depth for DLTS measurements is determined to be in the range of 150–250 nm. Notably, no correlation between tann and ND was observed for neither the H2 nor the Ar annealing. However, the as-received samples displayed a considerable spread in ND, and hence, a possible correlation between tann and ND might be masked. Previously, Polyakov et al. have shown that surface treatments with H-plasma lead to an increase in carrier concentration for EFG-grown bulk crystals with a (201) surface orientation, which the authors proposed to be related to the formation of shallow H-related donors.50 Interestingly, H-plasma treatments caused a decrease in carrier concentration for EFG-grown bulk crystals with a (010) surface orientation.38,50 This might be a result of distinct surface terminations on the (201) and (010) surfaces resulting from H treatment that influence the surface band bending.41 It has also been shown that H can contribute to the unintentional doping found in as-grown bulk crystals.40 

DLTS spectra recorded on as-received, 30 minH2-annealed, and 30 minAr-annealed crystals are presented in Fig. 2. The E1 peak (EA=0.60±0.04eV, σna6×1013cm2) is present in all three spectra (see the inset in Fig. 2 to discern the peak for the as-received and Ar-annealed sample). At temperatures of around 280 K, the onset of a signature commonly labeled as E2 can be seen, which has previously been shown to be related to substitutional Fe at tetrahedral Ga1 and octahedral Ga2 sites (FeGa1 and FeGa2).20,32,35 Notably, we did not observe the center commonly labeled as E2 after H2 anneals.11,23,32,33,36 For Ar anneals, however, a defect level around E2 appears as a small shoulder on the low-temperature side of E2 (its onset can be seen around 260 K in Fig. 2).

For the spectra presented in Fig. 2, the concentration of the E1 level is comparable for the as-received and Ar-annealed samples but considerably higher for the sample annealed in H2. Note that the peak position of the DLTS signature in the Ar-annealed sample is shifted to lower temperatures compared to that of the E1 level, which may indicate that Ar annealing results in the formation of other defect levels with a similar energy level position. Moreover, the DLTS signature of E1 (Fig. 2) is somewhat broader than that expected from a single level, as indicated by the simulated line. Thus, we cannot exclude that E1 consists of several overlapping levels. However, we were not able to resolve a finer structure in the E1 peak with the use of the high-resolution weighting function GS4.59 

FIG. 2.

DLTS spectra recorded on as-received, H2-, and Ar-annealed EFG-grown β-Ga2O3 bulk crystals. The rate window is (640ms)1. The data for the H2-annealed sample (circles) are modeled with a simulation (solid line). The H2 and Ar anneals were both performed at 900 °C for 30 min. The observed defect signatures are labeled. The inset shows the E1 peak for the as-received and Ar-annealed samples. The axes units of the inset are the same as for the main plot.

FIG. 2.

DLTS spectra recorded on as-received, H2-, and Ar-annealed EFG-grown β-Ga2O3 bulk crystals. The rate window is (640ms)1. The data for the H2-annealed sample (circles) are modeled with a simulation (solid line). The H2 and Ar anneals were both performed at 900 °C for 30 min. The observed defect signatures are labeled. The inset shows the E1 peak for the as-received and Ar-annealed samples. The axes units of the inset are the same as for the main plot.

Close modal

Figure 3 shows the E1 concentration in dependence of tann obtained from multiple DLTS measurements recorded on as-received, Ar-annealed, and H2-annealed samples. For the as-received and H2-annealed samples, the mean and standard deviation values are calculated from several diodes (between 3 and 14, the latter to check for lateral inhomogeneity) for the different tann. Note that the 15 and 60 min H2 annealing was performed solely on a single wafer, whereas the 75 min annealing was performed on a different wafer.

FIG. 3.

Mean and standard deviation values of the E1 concentration ([E1]) extracted from DLTS measurements on as-received, H2-, and Ar-annealed EFG-grown bulk crystals in dependence of tann. The λ-correction was taken into account for computing [E1].

FIG. 3.

Mean and standard deviation values of the E1 concentration ([E1]) extracted from DLTS measurements on as-received, H2-, and Ar-annealed EFG-grown bulk crystals in dependence of tann. The λ-correction was taken into account for computing [E1].

Close modal

From Fig. 3, one can observe that the mean E1 concentration in as-received bulk crystals is low. Indeed, for some of the diodes on the as-received bulk crystals, the E1 concentration was below the detection limit of around 5 × 1012 cm−3 and thus not observed in the DLTS measurements. The diodes on the as-received samples that displayed the presence of E1 had an Nt of around 5 × 1013 cm−3 taking the λ-correction into account. The samples annealed in argon similarly displayed a low E1 concentration of around 1×1014 cm−3. Moreover, the Ar-annealed samples do not show a systematic increase in the E1 concentration with increasing tann. For the Ar-annealed samples, it should be noted that the E1 concentration was extracted treating the shifted peaks as pertaining to E1, and hence, the calculated concentrations can be considered an upper bound. The samples annealed in H2, in contrast, display a considerably larger concentration of E1 in the range of 1×1015 cm−3. For annealing times up to and including 60 min, the mean E1 concentration also increases with increased time.

For the sample annealed in H2 for 75 min, a slightly lower concentration of E1 is measured compared to that of the 60 min ones but still substantially above that of the as-received. This may indicate that for long annealing times, multiple defect reactions may influence the overall E1 concentration. In addition, as the 60 min-annealed and 75 min-annealed diodes stem from two separate wafers, initial differences in the relative and absolute defect concentrations can affect the resulting E1 concentration. Nevertheless, the results displayed in Figs. 2 and 3 lead us to propose that E1 is associated with a H-related defect.

Probing the thermal and field dependent stability of defect levels can provide valuable information for the identification of defects. Several defect levels in β-Ga2O3 have previously been shown to be metastable36,49 with the use of RBA and ZBA. For example, we have previously found that E2 formed by H implantation can be reversibly introduced and removed by performing RBA and ZBA, respectively, at temperatures of around 650 K.36Figure 4 shows the results for the E1 level following RBA and ZBA cycles. More specifically, DLTS spectra recorded after 60 minH2 annealing and after subsequent RBA and ZBA at 650 K are presented. A notable finding is that the peak intensity shows an insignificant change after the annealing cycles. The unchanged intensity suggests that E1 is related to a stable defect. We observe only a slight increase (decrease) in the peak intensity (temperature position) of the E2 signature, following RBA and no further change following ZBA. Furthermore, no distinct shoulder, which would correspond to E2, emerges on the low-temperature side of E2 after RBA at 650 K.

FIG. 4.

DLTS spectra recorded on a sample annealed in H2 for a duration of 60 min. The initial DLTS spectra and the DLTS spectra following RBA and ZBA at 650 K are shown. The rate window is (640ms)1.

FIG. 4.

DLTS spectra recorded on a sample annealed in H2 for a duration of 60 min. The initial DLTS spectra and the DLTS spectra following RBA and ZBA at 650 K are shown. The rate window is (640ms)1.

Close modal

Hybrid-functional calculations were performed to explore potential defects that might give rise to the E1 center, assuming a H-related origin. Previous calculations indicate that Hi behaves exclusively as a shallow donor in β-Ga2O3, as the predicted ε(+/) level is close to EC.45 Isolated H could also occur in the form of interstitial molecular hydrogen (H2)i, which we find to be electrically inactive and stable only in n-type material with a maximum binding energy of 0.85 eV. However, Hi and (H2)i are expected to be highly mobile.45 Indeed, using the climbing-image nudged elastic band method73 and the strongly constrained and appropriately normed semilocal functional,74 we calculate migration barriers along the b axis of 0.24 and 0.61 eV for Hi and (H2)i, respectively. For this reason, H most likely occurs in a trapped form, such as a defect complex involving an intrinsic defect or possibly another impurity; Si, Al, Fe, and Ir are commonly found in EFG-grown β-Ga2O3 crystals.32,75

Figures 5(a) and 5(b) show the formation and binding energy diagrams, respectively, of various H-related defect complexes exhibiting charge-state transition levels in the vicinity of E1, as discussed below. The formation energies of other defects mentioned below are reported elsewhere.33,45,76,77 The notation used for the defects is in accordance with Ref. 77.

FIG. 5.

(a) Calculated formation energies under (left) Ga-rich and (right) O-rich conditions of various H-related defect complexes in β-Ga2O3. These defects display thermodynamic charge-state transition levels that are close in Fermi-level position to the measured EA for E1 (gray vertical bar 0.6 eV ± 0.1 eV below EC). Note that all corresponding transitions show negative-U behavior. The inset shows the /2 transitions for the singly hydrogenated Ga–O divacancies. The axes units of the inset are the same as for the main plot. (b) Calculated binding energies for the H-related complexes.

FIG. 5.

(a) Calculated formation energies under (left) Ga-rich and (right) O-rich conditions of various H-related defect complexes in β-Ga2O3. These defects display thermodynamic charge-state transition levels that are close in Fermi-level position to the measured EA for E1 (gray vertical bar 0.6 eV ± 0.1 eV below EC). Note that all corresponding transitions show negative-U behavior. The inset shows the /2 transitions for the singly hydrogenated Ga–O divacancies. The axes units of the inset are the same as for the main plot. (b) Calculated binding energies for the H-related complexes.

Close modal

We start by considering intrinsic defects that can trap H. Previous calculations have shown that H can be trapped by VO, resulting in a HO complex that behaves as a shallow donor.45 Interestingly, we find that the HO2 configuration can be stabilized also in the single-negative charge state for Fermi-level positions close to EC (see Fig. 5). As shown in Fig. 6, the single-negative charge state involves a large structural rearrangement, where H moves off the vacant threefold coordinated O2 site to form a bond with the adjacent Ga2 atom, and two electrons are captured in a localized state. Note that charge-neutral HO2 is unstable for any position of the Fermi level, resulting in negative-U behavior;78 i.e., the thermodynamic charge-state transition occurs directly from single positive to single negative in Fig. 5(a). For a negative-U center, the peak observed in a conventional DLTS spectrum will correspond to the emission of two electrons, but EA will be determined by the first electron emission, corresponding to the ε(0/) level for HO2.79 

FIG. 6.

Relaxed structures of the H-related defect complexes discussed as potential E1 origins. The transition from the + to the charge state of HO2 corresponds to H moving off the vacant O2 site to form a bond with the adjacent Ga2 atom, allowing two electrons to be trapped in a localized defect state (blue isosurface). For the (SiGa1H)0 and (SnGa2H)0 complexes, the arrows indicate the O site where an O–H bond is formed for their 2+ charge states in Fig. 5.

FIG. 6.

Relaxed structures of the H-related defect complexes discussed as potential E1 origins. The transition from the + to the charge state of HO2 corresponds to H moving off the vacant O2 site to form a bond with the adjacent Ga2 atom, allowing two electrons to be trapped in a localized defect state (blue isosurface). For the (SiGa1H)0 and (SnGa2H)0 complexes, the arrows indicate the O site where an O–H bond is formed for their 2+ charge states in Fig. 5.

Close modal

As shown in Table I, the corresponding Ei value for HO2 is 0.68 eV with a small Eb of 0.10 eV, which is close to the measured EA of 0.6 eV for E1. However, HO2 is expected to show a low thermal stability, with a maximum binding energy of 0.69 eV when the Fermi-level position is near EC, as shown in Fig. 5(b). Combining this with the low Hi migration barrier, one would thus expect HO2 to dissociate relatively easily.33 This is hard to reconcile with the apparent stability of E1 upon ZBA and RBA up to 650 K.

TABLE I.

Calculated CC model parameters for the process of nonradiative emission of an electron from different defect complexes involving H to EC, including Ei, Eb, dFCg, dFCe, and ΔQ.

Defect and transitionEi (eV)Eb (eV)dFCg/e (eV)ΔQ (amu1/2Å)
HO2 (0/−) 0.68 0.10 1.45 / 1.53 2.78 
VGaibH–VO1 (−/2−) 0.52 0.19 1.48 / 2.01 5.59 
VGa1H–VO1 (−/2−) 0.49 0.20 1.46 / 2.03 6.41 
SiGa1H (+/0) ≤0.99 … … … 
SnGa2H (+/0) ≤1.07 … … … 
Defect and transitionEi (eV)Eb (eV)dFCg/e (eV)ΔQ (amu1/2Å)
HO2 (0/−) 0.68 0.10 1.45 / 1.53 2.78 
VGaibH–VO1 (−/2−) 0.52 0.19 1.48 / 2.01 5.59 
VGa1H–VO1 (−/2−) 0.49 0.20 1.46 / 2.03 6.41 
SiGa1H (+/0) ≤0.99 … … … 
SnGa2H (+/0) ≤1.07 … … … 

As a donor, Hi is particularly likely to form stable complexes with acceptors, such as VGa.33,80 Indeed, the main O–H vibrational line observed by FT-IR in the hydrogen-annealed material is caused by VGaib2H.42 However, the calculated thermodynamic charge-state transition levels of VGa are located in excess of 1.8 eV below EC, and complexing with H tends to shift these levels to even lower Fermi-level positions.33 

Another possibility is Ga–O divacancies (VGaVO), which exhibit negative-Uε(/3) levels in the upper part of the bandgap that are associated with the formation of Ga–Ga dimers at VO.33,77VGaVO can occur in a large number of crystallographically inequivalent configurations. However, the negative-U charge-state transition levels tend to (i) cluster within narrow ranges of Fermi-level positions, depending on the combination of tetrahedral Ga1 and octahedral Ga2 in the dimer, and (ii) shift to lower Fermi-level positions when VGaVO is hydrogenated.77 We have previously discussed certain configurations of the isolated and doubly hydrogenated VGaVO as potential defect origins of the E2 center.77 As shown in Fig. 5, we find that the singly hydrogenated divacancies VGaibH–VO1 and VGa1H–VO1 exhibit ε(0/2) levels that are close to EC and E1 (four additional configurations with similar level positions can be found in Ref. 77, but these are not included here as they are higher in energy). As shown in Fig. 6, the 2- charge state correspond to the formation of a Ga2–Ga2 dimer at VO1. In this case, ε(/2) is the relevant level for comparison with DLTS, and the corresponding Ei (Eb) values are 0.52 (0.19) and 0.49 (0.20) eV for VGaibH–VO1 and VGa1H–VO1, respectively. These energies are also compatible with E1.

In contrast to the HO2 complex, VGaibH–VO1 and VGa1H–VO1 are expected to show high thermal stability, with binding energies in excess of 2.2 eV under n-type conditions, as shown in Fig. 5(b). These binding energies are comparable to the 2.62 eV predicted for VGaib2H.80 Furthermore, the possibility of E1 being composed of several overlapping defects is in agreement with this defect model, as there are several configurations of the singly hydrogenated divacancy exhibiting similar activation energies.77 

It should be noted that the VGa2H–VO2 and VGaiaH–VO2b configurations, which have the ε(0/2) level at 1.7 eV below EC, are 0.4 eV lower in formation energy than VGaibH–VO1 when EF=EC.77VGaibH–VO1 becomes preferred when the Fermi-level position is below EC0.46 eV.77 Moreover, a second H can be trapped by VGaibH–VO1, resulting in the VGaibH–HO1 configuration when EF=EC. The VGaibH–HO1 complex has a H binding energy of 2.34 eV and does not exhibit any levels in the vicinity of E1. For these reasons, the concentration of VGaibH–VO1 (and other E1 compatible configurations), relative to divacancies in other configurations and with different numbers of H, can be expected to depend on several factors, such as the concentration ratio of H to VGaVO and other traps, the Fermi-level position, the temperature, and the energy barriers associated with transformations between different isolated and hydrogenated VGaVO configurations.77,81

Finally, we consider complexes between H and other common impurities in β-Ga2O3. An obvious candidate is the FeGa acceptor (E2), which displays ε(0/) levels at 0.6-0.7 eV below EC.20,32,35 However, Varley76 calculated a binding energy of 0.7 eV for the FeGa2H complex and a ε(+/0) level located 1.3 eV below EC.76 Similarly, polaronic acceptor impurities, such as Mg, exhibit even deeper levels that are also shifted to lower Fermi-level positions when hydrogenated.82,83 Polyakov et al.38 have suggested that H can passivate shallow donor impurities by forming charge-neutral complexes. We have investigated this possibility for silicon and tin donors in their most favorable configuration, SiGa1 and SnGa2 (Sn is a commonly used n-type dopant and is included for comparison). Although cationic Hi+ is the energetically preferred form of isolated H, we indeed find that anionic Hi can be stabilized in the vicinity of these Ga substitutional shallow donor impurities, resulting in charge-neutral SiGa1H and SnGa2H pairs under n-type conditions, as shown in Fig. 5(a). However, formation of these complexes might be suppressed by screened Coulomb repulsion, as both constituents are positively charged for any position of the Fermi level in the bandgap. Moreover, these complexes are only stable in n-type material, with modest binding energies of up to 0.58 and 0.51 eV for SiGa1H and SnGa2H, respectively, as shown in Fig. 5(b). Additionally, their stability rapidly decreases with the Fermi level, indicating likely complex dissociation under RBA conditions. For lower Fermi-level positions, Hi+ is preferred over Hi, also in the vicinity of the donor. Indeed, the 2+ charge states shown in Fig. 5 correspond to a Hi+ immediately adjacent to SiGa+ or SnGa+ (H forms a bond with the O indicated by arrows in Fig. 6). The single-positive charge state of the complex is similar but has an electron occupying a delocalized perturbed host state66 just below EC. This prevents an accurate evaluation of ε(+/0), as the 160-atom supercell is not sufficiently large for such spatially extended defect states.84 The upper estimates of 1 eV in Table I assume a donor ionization energy of zero, which corresponds to the ε(2+/+) level being located at EC. Nonetheless, the low thermal and bias-induced stability of these donor complexes are incompatible with those observed for E1.

In summary, we have investigated the influence of H2 and Ar annealing of n-type EFG-grown β-Ga2O3 on the presence of the E1 center. Using FT-IR, we confirmed that H is incorporated into the bulk crystals during H2 heat treatments. Notably, the H2 heat treatments did not lead to any considerable changes in charge-carrier concentration. Using DLTS, it was shown that the H2 annealing promotes the defect level E1, suggesting a H-related defect origin for the E1 center. Based on comparison with hybrid-functional calculations for defect complexes involving H, specific defect origins of E1 are discussed. We find three different types of H-related defects exhibiting charge-state transition levels and capture barriers that are compatible with the measured activation energy for E1, including (i) the ε(+/) level of HO2, (ii) the ε(/2) level of singly hydrogenated Ga–O divacancies exhibiting Ga2–Ga2 dimers (VGaibH–VO1 and VGa1H–VO1 being the lowest energy configurations),77 and (iii) the ε(+/0) level of Ga substitutional shallow donor impurities passivated by anionic Hi, e.g., SiGa1H and SnGa2H. Among these defect candidates, only (ii) is consistent with the apparent stability of E1 upon RBA and ZBA up to 650 K.

Financial support is acknowledged from the Research Council of Norway through the GO2DEVICE project (Project No. 301740), the FUNDAMeNT project (Project No. 251131), the GO-POW project (Project No. 314017), the Norwegian Micro- and Nano-Fabrication Facility (NorFab, Project No. 295864), and the Faculty of Mathematics and Natural Sciences at the University of Oslo via the strategic research initiative FOXHOUND. Funding for this work was also provided by the Norwegian Research Council through the Research Center for Sustainable Solar Cell Technology (FME SUSOLTECH, Project No. 257639). This work was partially performed under the auspices of the U.S. DOE by the Lawrence Livermore National Laboratory (LLNL) under Contract No. DE-AC52-07NA27344 and partially supported by the LLNL Laboratory Directed Research and Development funding under Project No. 22-SI-003 and by the Critical Materials Institute, an Energy Innovation Hub funded by the U.S. DOE, Office of Energy Efficiency and Renewable Energy, Advanced Manufacturing Office.

The authors have no conflicts to disclose.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
C.
Janowitz
,
V.
Scherer
,
M.
Mohamed
,
A.
Krapf
,
H.
Dwelk
,
R.
Manzke
,
Z.
Galazka
,
R.
Uecker
,
K.
Irmscher
, and
R.
Fornari
et al.,
New J. Phys.
13
,
085014
(
2011
).
2.
T.
Onuma
,
S.
Saito
,
K.
Sasaki
,
K.
Goto
,
T.
Masui
,
T.
Yamaguchi
,
T.
Honda
,
A.
Kuramata
, and
M.
Higashiwaki
,
Appl. Phys. Lett.
108
,
101904
(
2016
).
3.
T.
Onuma
,
S.
Saito
,
K.
Sasaki
,
T.
Masui
,
T.
Yamaguchi
,
T.
Honda
, and
M.
Higashiwaki
,
Jpn. J. Appl. Phys.
54
,
112601
(
2015
).
4.
M.
Higashiwaki
,
K.
Sasaki
,
H.
Murakami
,
Y.
Kumagai
,
A.
Koukitu
,
A.
Kuramata
,
T.
Masui
, and
S.
Yamakoshi
,
Semicond. Sci. Technol.
31
,
034001
(
2016
).
5.
S.
Nakagomi
,
T.
Momo
,
S.
Takahashi
, and
Y.
Kokubun
,
Appl. Phys. Lett.
103
,
072105
(
2013
).
6.
S.
Nakagomi
,
T.
Sato
,
Y.
Takahashi
, and
Y.
Kokubun
,
Sens. Actuator A Phys.
232
,
208
(
2015
).
7.
M. A.
Mastro
,
A.
Kuramata
,
J.
Calkins
,
J.
Kim
,
F.
Ren
, and
S.
Pearton
,
ECS J. Solid State Sci. Technol.
6
,
P356
(
2017
).
8.
Z.
Galazka
,
Semicond. Sci. Technol.
33
,
113001
(
2018
).
9.
S. J.
Pearton
,
J.
Yang
,
P. H.
Cary
,
F.
Ren
,
J.
Kim
,
M. J.
Tadjer
, and
M. A.
Mastro
,
Appl. Phys. Rev.
5
,
011301
(
2018
).
10.
P.
Blood
and
J.
Orton
,
The Electrical Characterization of Semiconductors: Majority Carriers and Electron States
(
Academic Press
,
1992
).
11.
J. F.
Mcglone
,
Z.
Xia
,
Y.
Zhang
,
C.
Joishi
,
S.
Lodha
,
S.
Rajan
,
S. A.
Ringel
, and
A. R.
Arehart
,
IEEE Electron Device Lett.
39
,
1042
(
2018
).
12.
J. F.
McGlone
,
Z.
Xia
,
C.
Joishi
,
S.
Lodha
,
S.
Rajan
,
S.
Ringel
, and
A. R.
Arehart
,
Appl. Phys. Lett.
115
,
153501
(
2019
).
13.
S.-S.
Huang
,
R.
Lopez
,
S.
Paul
,
A. T.
Neal
,
S.
Mou
,
M.-P.
Houng
, and
J. V.
Li
,
Jpn. J. Appl. Phys.
57
,
091101
(
2018
).
14.
L.
Huang
,
Q.
Feng
,
G.
Han
,
F.
Li
,
X.
Li
,
L.
Fang
,
X.
Xing
,
J.
Zhang
, and
Y.
Hao
,
IEEE Photon. J.
9
,
6803708
(
2017
).
15.
S.
Bhandari
and
M. E.
Zvanut
,
J. Appl. Phys.
127
,
065704
(
2020
).
16.
S.
Bhandari
,
M. E.
Zvanut
, and
J. B.
Varley
,
J. Appl. Phys.
126
,
165703
(
2019
).
17.
C. A.
Lenyk
,
T. D.
Gustafson
,
L. E.
Halliburton
, and
N. C.
Giles
,
J. Appl. Phys.
126
,
245701
(
2019
).
18.
H.
Gao
,
S.
Muralidharan
,
N.
Pronin
,
M. R.
Karim
,
S. M.
White
,
T.
Asel
,
G.
Foster
,
S.
Krishnamoorthy
,
S.
Rajan
,
L. R.
Cao
,
M.
Higashiwaki
,
H.
von Wenckstern
,
M.
Grundmann
,
H.
Zhao
,
D. C.
Look
, and
L. J.
Brillson
,
Appl. Phys. Lett.
112
,
242102
(
2018
).
19.
H.
Gao
,
S.
Muralidharan
,
M. R.
Karim
,
S. M.
White
,
L. R.
Cao
,
K.
Leedy
,
H.
Zhao
,
D. C.
Look
, and
L. J.
Brillson
,
J. Phys. D: Appl. Phys.
53
,
465102
(
2020
).
20.
C.
Zimmermann
,
Y. K.
Frodason
,
V.
Rønning
,
J. B.
Varley
, and
L.
Vines
,
New J. Phys.
22
,
063033
(
2020
).
21.
C.
Zimmermann
,
V.
Rønning
,
Y. K.
Frodason
,
V.
Bobal
,
L.
Vines
, and
J. B.
Varley
,
Phys. Rev. Mater.
4
,
074605
(
2020
).
22.
E.
Farzana
,
E.
Ahmadi
,
J. S.
Speck
,
A. R.
Arehart
, and
S. A.
Ringel
,
J. Appl. Phys.
123
,
161410
(
2018
).
23.
A. Y.
Polyakov
,
N. B.
Smirnov
,
I. V.
Shchemerov
,
E. B.
Yakimov
,
J.
Yang
,
F.
Ren
,
G.
Yang
,
J.
Kim
,
A.
Kuramata
, and
S. J.
Pearton
,
Appl. Phys. Lett.
112
,
032107
(
2018
).
24.
A. Y.
Polyakov
,
N. B.
Smirnov
,
I. V.
Shchemerov
,
E. B.
Yakimov
,
S. J.
Pearton
,
C.
Fares
,
J.
Yang
,
F.
Ren
,
J.
Kim
,
P. B.
Lagov
,
V. S.
Stolbunov
, and
A.
Kochkova
,
Appl. Phys. Lett.
113
,
092102
(
2018
).
25.
Z.
Zhang
,
E.
Farzana
,
A. R.
Arehart
, and
S. A.
Ringel
,
Appl. Phys. Lett.
108
,
052105
(
2016
).
26.
A. Y.
Polyakov
,
N. B.
Smirnov
,
I. V.
Shchemerov
,
D.
Gogova
,
S. A.
Tarelkin
, and
S. J.
Pearton
,
J. Appl. Phys.
123
,
115702
(
2018
).
27.
A. Y.
Polyakov
,
N. B.
Smirnov
,
I. V.
Shchemerov
,
S. J.
Pearton
,
F.
Ren
,
A. V.
Chernykh
, and
A. I.
Kochkova
,
Appl. Phys. Lett.
113
,
142102
(
2018
).
28.
E.
Farzana
,
M. F.
Chaiken
,
T. E.
Blue
,
A. R.
Arehart
, and
S. A.
Ringel
,
APL Mater.
7
,
022502
(
2019
).
29.
Y.
Nakano
,
ECS J. Solid State Sci. Technol.
6
,
P615
(
2017
).
30.
H.
Ghadi
,
J. F.
McGlone
,
C. M.
Jackson
,
E.
Farzana
,
Z.
Feng
,
A. F. M.
A. U. Bhuiyan
,
H.
Zhao
,
A. R.
Arehart
, and
S. A.
Ringel
,
APL Mater.
8
,
021111
(
2020
).
31.
K.
Irmscher
,
Z.
Galazka
,
M.
Pietsch
,
R.
Uecker
, and
R.
Fornari
,
J. Appl. Phys.
110
,
063720
(
2011
).
32.
M. E.
Ingebrigtsen
,
J. B.
Varley
,
A. Y.
Kuznetsov
,
B. G.
Svensson
,
G.
Alfieri
,
A.
Mihaila
,
U.
Badstübner
, and
L.
Vines
,
Appl. Phys. Lett.
112
,
042104
(
2018
).
33.
M. E.
Ingebrigtsen
,
A. Y.
Kuznetsov
,
B. G.
Svensson
,
G.
Alfieri
,
A.
Mihaila
,
U.
Badstübner
,
A.
Perron
,
L.
Vines
, and
J. B.
Varley
,
APL Mater.
7
,
022510
(
2019
).
34.
A. Y.
Polyakov
,
N. B.
Smirnov
,
I. V.
Shchemerov
,
S. J.
Pearton
,
F.
Ren
,
A. V.
Chernykh
,
P. B.
Lagov
, and
T. V.
Kulevoy
,
APL Mater.
6
,
096102
(
2018
).
35.
C.
Zimmermann
,
Y. K.
Frodason
,
A. W.
Barnard
,
J. B.
Varley
,
K.
Irmscher
,
Z.
Galazka
,
A.
Karjalainen
,
W. E.
Meyer
,
F. D.
Auret
, and
L.
Vines
,
Appl. Phys. Lett.
116
,
072101
(
2020
).
36.
C.
Zimmermann
,
E. F.
Verhoeven
,
Y. K.
Frodason
,
P. M.
Weiser
,
J. B.
Varley
, and
L.
Vines
,
J. Phys. D: Appl. Phys.
53
,
464001
(
2020
).
37.
M. M.
Islam
,
M. O.
Liedke
,
D.
Winarski
,
M.
Butterling
,
A.
Wagner
,
P.
Hosemann
,
Y.
Wang
,
B.
Uberuaga
, and
F. A.
Selim
,
Sci. Rep.
10
,
6134
(
2020
).
38.
A. Y.
Polyakov
,
I.-H.
Lee
,
N. B.
Smirnov
,
E. B.
Yakimov
,
I. V.
Shchemerov
,
A. V.
Chernykh
,
A. I.
Kochkova
,
A. A.
Vasilev
,
F.
Ren
,
P. H.
Carey
, and
S. J.
Pearton
,
Appl. Phys. Lett.
115
,
032101
(
2019
).
39.
Y.
Qin
,
M.
Stavola
,
W. B.
Fowler
,
P.
Weiser
, and
S. J.
Pearton
,
ECS J. Solid. State. Sci.
8
,
Q3103
(
2019
).
40.
Z.
Galazka
,
K.
Irmscher
,
R.
Schewski
,
I. M.
Hanke
,
M.
Pietsch
,
S.
Ganschow
,
D.
Klimm
,
A.
Dittmar
,
A.
Fiedler
,
T.
Schroeder
, and
M.
Bickermann
,
J. Cryst. Growth
529
,
125297
(
2020
).
41.
J. E. N.
Swallow
,
J. B.
Varley
,
L. A. H.
Jones
,
J. T.
Gibbon
,
L. F. J.
Piper
,
V. R.
Dhanak
, and
T. D.
Veal
,
APL Mater.
7
,
022528
(
2019
).
42.
P.
Weiser
,
M.
Stavola
,
W. B.
Fowler
,
Y.
Qin
, and
S.
Pearton
,
Appl. Phys. Lett.
112
,
232104
(
2018
).
43.
J. B.
Varley
,
H.
Peelaers
,
A.
Janotti
, and
C. G. V.
de Walle
,
J. Phys.: Condens. Matter
23
,
334212
(
2011
).
44.
P. D. C.
King
,
I.
McKenzie
, and
T. D.
Veal
,
Appl. Phys. Lett.
96
,
062110
(
2010
).
45.
J. B.
Varley
,
J. R.
Weber
,
A.
Janotti
, and
C. G. V.
de Walle
,
Appl. Phys. Lett.
97
,
142106
(
2010
).
46.
J. R.
Ritter
,
J.
Huso
,
P. T.
Dickens
,
J. B.
Varley
,
K. G.
Lynn
, and
M. D.
McCluskey
,
Appl. Phys. Lett.
113
,
052101
(
2018
).
47.
C.
Pansegrau
,
J.
Jesenovec
,
J. S.
McCloy
, and
M. D.
McCluskey
,
Appl. Phys. Lett.
119
,
102104
(
2021
).
48.
M.
Higashiwaki
and
S.
Fujita
,
Gallium Oxide
(
Springer International Publishing
,
2020
).
49.
M. E.
Ingebrigtsen
,
A. Y.
Kuznetsov
,
B. G.
Svensson
,
G.
Alfieri
,
A.
Mihaila
, and
L.
Vines
,
J. Appl. Phys.
125
,
185706
(
2019
).
50.
A. Y.
Polyakov
,
I.-H.
Lee
,
A.
Miakonkikh
,
A. V.
Chernykh
,
N. B.
Smirnov
,
I. V.
Shchemerov
,
A. I.
Kochkova
,
A. A.
Vasilev
, and
S. J.
Pearton
,
J. Appl. Phys.
127
,
175702
(
2020
).
51.
A. Y.
Polyakov
,
N. B.
Smirnov
,
I. V.
Shchemerov
,
A. A.
Vasilev
,
E. B.
Yakimov
,
A. V.
Chernykh
,
A. I.
Kochkova
,
P. B.
Lagov
,
Y. S.
Pavlov
,
O. F.
Kukharchuk
,
A. A.
Suvorov
,
N. S.
Garanin
,
I.-H.
Lee
,
M.
Xian
,
F.
Ren
, and
S. J.
Pearton
,
J. Phys. D: Appl. Phys.
53
,
274001
(
2020
).
52.
H.
Aida
,
K.
Nishiguchi
,
H.
Takeda
,
N.
Aota
,
K.
Sunakawa
, and
Y.
Yaguchi
,
Jpn. J. Appl. Phys.
47
,
8506
(
2008
).
53.
A.
Kuramata
,
K.
Koshi
,
S.
Watanabe
,
Y.
Yamaoka
,
T.
Masui
, and
S.
Yamakoshi
,
Jpn. J. Appl. Phys.
55
,
1202A2
(
2016
).
54.
See https://www.tamuracorp.com/products/gao/index.html for “Tamura Corporation: β-Ga2O3 Crystals Grown by Edge-Defined Film-Fed Growth” (2020); accessed 12 May 2020.
55.
M.
Stavola
and
W. B.
Fowler
,
J. Appl. Phys.
123
,
161561
(
2018
).
56.
M. E.
Ingebrigtsen
,
L.
Vines
,
G.
Alfieri
,
A.
Mihaila
,
U.
Badstübner
,
B. G.
Svensson
, and
A.
Kuznetsov
,
Mater. Sci. Forum
897
,
755
(
2017
).
57.
M.
Passlack
,
N. E. J.
Hunt
,
E. F.
Schubert
,
G. J.
Zydzik
,
M.
Hong
,
J. P.
Mannaerts
,
R. L.
Opila
, and
R. J.
Fischer
,
Appl. Phys. Lett.
64
,
2715
(
1994
).
58.
B. G.
Svensson
,
K.-H.
Rydén
, and
B. M. S.
Lewerentz
,
J. Appl. Phys.
66
,
1699
(
1989
).
59.
A. A.
Istratov
,
J. Appl. Phys.
82
,
2965
(
1997
).
60.
D. V.
Lang
,
J. Appl. Phys.
45
,
3023
(
1974
).
61.
P. E.
Blöchl
,
Phys. Rev. B
50
,
17953
(
1994
).
62.
G.
Kresse
and
D.
Joubert
,
Phys. Rev. B
59
,
1758
(
1999
).
63.
A. V.
Krukau
,
O. A.
Vydrov
,
A. F.
Izmaylov
, and
G. E.
Scuseria
,
J. Chem. Phys.
125
,
224106
(
2006
).
64.
G.
Kresse
and
J.
Furthmüller
,
Phys. Rev. B
54
,
11169
(
1996
).
65.
J.
Åhman
,
G.
Svensson
, and
J.
Albertsson
,
Acta Crystallogr. C Cryst. Struct. Commun.
52
,
1336
(
1996
).
66.
C.
Freysoldt
,
B.
Grabowski
,
T.
Hickel
,
J.
Neugebauer
,
G.
Kresse
,
A.
Janotti
, and
C. G. V.
de Walle
,
Rev. Mod. Phys.
86
,
253
(
2014
).
67.
C.
Freysoldt
,
J.
Neugebauer
, and
C. G. V.
de Walle
,
Phys. Rev. Lett.
102
,
016402
(
2009
).
68.
Y.
Kumagai
and
F.
Oba
,
Phys. Rev. B
89
,
195205
(
2014
).
69.
T.
Gake
,
Y.
Kumagai
,
C.
Freysoldt
, and
F.
Oba
,
Phys. Rev. B
101
,
020102
(
2020
).
70.
A. M.
Stoneham
,
Theory of Defects in Solids
(
Oxford University
,
1975
).
71.
D.
Wickramaratne
,
C. E.
Dreyer
,
B.
Monserrat
,
J.-X.
Shen
,
J. L.
Lyons
,
A.
Alkauskas
, and
C. G. V.
de Walle
,
Appl. Phys. Lett.
113
,
192106
(
2018
).
72.
A.
Alkauskas
,
M. D.
McCluskey
, and
C. G. V.
de Walle
,
J. Appl. Phys.
119
,
181101
(
2016
).
73.
G.
Henkelman
,
B. P.
Uberuaga
, and
H.
Jansson
,
J. Chem. Phys.
113
,
9901
(
2000
).
74.
J.
Sun
,
A.
Ruzsinszky
, and
J.
Perdew
,
Phys. Rev. Lett.
115
,
036402
(
2015
).
75.
A.
Kuramata
,
K.
Koshi
,
S.
Watanabe
, and
Y.
Yamaoka
, “Floating zone method, edge-defined film-fed growth method, and wafer manufacturing,” in Gallium Oxide: Materials Properties, Crystal Growth, and Devices, edited by M. Higashiwaki and S. Fujita (Springer International Publishing, Cham, 2020), pp. 57–75.
76.
J. B.
Varley
, “First-principles calculations 2,” in Gallium Oxide: Materials Properties, Crystal Growth, and Devices, edited by M. Higashiwaki and S. Fujita (Springer International Publishing, Cham, 2020), pp. 329–348.
77.
Y. K.
Frodason
,
C.
Zimmermann
,
E. F.
Verhoeven
,
P. M.
Weiser
,
L.
Vines
, and
J. B.
Varley
,
Phys. Rev. Mater.
5
,
025402
(
2021
).
78.
P. W.
Anderson
,
Phys. Rev. Lett.
34
,
953
(
1975
).
79.
C. G.
Hemmingsson
,
N. T.
Son
,
A.
Ellison
,
J.
Zhang
, and
E.
Janzén
,
Phys. Rev. B
58
,
R10119
(
1998
).
80.
V. M.
Reinertsen
,
P. M.
Weiser
,
Y. K.
Frodason
,
M. E.
Bathen
,
L.
Vines
, and
K. M.
Johansen
,
Appl. Phys. Lett.
117
,
232106
(
2020
).
81.
W. B.
Fowler
,
M.
Stavola
,
Y.
Qin
, and
P.
Weiser
,
Appl. Phys. Lett.
117
,
142101
(
2020
).
82.
J. L.
Lyons
,
Semicond. Sci. Technol.
33
,
05LT02
(
2018
).
83.
Y. K.
Frodason
,
K. M.
Johansen
,
L.
Vines
, and
J. B.
Varley
,
J. Appl. Phys.
127
,
075701
(
2020
).
84.
M. W.
Swift
,
H.
Peelaers
,
S.
Mu
,
J. J. L.
Morton
, and
C. G. V.
de Walle
,
npj Computat. Mater.
6
,
181
(
2020
).