The ability to achieve highly resistive beta-phase gallium oxide (β-Ga2O3) layers and substrates is critical for β-Ga2O3 high voltage and RF devices. To date, the most common approach involves doping with iron (Fe), which generates a moderately deep acceptor-like defect state located at EC-0.8 eV in the β-Ga2O3 bandgap. Recently, there has been growing interest in alternative acceptors, such as magnesium (Mg) and nitrogen (N), due to their predicted deeper energy levels, which could avoid inadvertent charge modulation during device operation. In this work, a systematic study that makes direct correlations between the introduction of N using ion implantation and the observation of a newly observed deep level at EC-2.9 eV detected by deep-level optical spectroscopy (DLOS) is presented. The concentration of this state displayed a monotonic dependence with N concentration over a range of implant conditions, as confirmed by secondary ion mass spectrometry (SIMS). With a near 1:1 match in absolute N and EC-2.9 eV trap concentrations from SIMS and DLOS, respectively, which also matched the measured removal of free electrons from capacitance-voltage studies, this indicates that N contributes a very efficiently incorporated compensating defect. Density functional theory calculations confirm the assignment of this state to be an N (0/−1) acceptor with a configuration of N occupying the oxygen site III [NO(III)]. The near ideal efficiency for this state to compensate free electrons and its location toward the midgap region of the β-Ga2O3 bandgap demonstrates the potential of N doping as a promising approach for producing semi-insulating β-Ga2O3.

Beta-phase gallium oxide (β-Ga2O3) possesses an ultra-wide bandgap (UWBG) of 4.6–4.8 eV,1–3 inherent n-type conductivity with donor impurities,4,5 and projected high radiation hardness.6,7 The advantageous material properties of β-Ga2O3 lead to outstanding device figures of merit projections that can revolutionize high voltage and RF devices. As a result of these anticipated performance specifications, a large body of research has focused on advancing the quality of gallium oxide materials, heterostructures, and a wide variety of devices. To date, high breakdown β-Ga2O3 metal-semiconductor field-effect transistors (MESFETs),8 high 2DEG charge densities in (AlxGa1−x)2O3/Ga2O3 modulation-doped field-effect transistors (MODFETS),9 and enhancement mode β-Ga2O3 transistors10 have been epitaxially grown on Fe-doped β-Ga2O3 semi-insulating substrates. For these lateral devices, it is necessary to have semi-insulating or highly resistive buffers or substrates acting as a blocking layer to prevent buffer leakage. Regardless of the growth method, the unintentional n-type conductivity can be compensated by introducing deep acceptor-like impurities. Density functional theory (DFT) calculations have identified several impurities that are predicted to form deep acceptor-like bandgap states, such as iron (Fe), nitrogen (N), cobalt (Co), and magnesium (Mg).11–14 These deep acceptors can be used as dopants to compensate for excess electrons and create highly resistive layers. Of these, Fe has been adopted in most cases for achieving highly resistive and semi-insulating bulk β-Ga2O3 substrates and epitaxial layers.12,15,16

Due to this popularity, much work has been published regarding the properties of Fe in β-Ga2O3. Computational results validated by deep-level (thermal) transient spectroscopy (DLTS), electron paramagnetic resonance (EPR), and cathodoluminescence (CL) experiments have demonstrated that Fe incorporates into n-type β-Ga2O3 as an acceptor-like defect.12,17–26 It is predicted to preferentially substitute on the Ga sites in the β-Ga2O3 lattice, creating a bandgap state at EC-0.8 eV (commonly referred to as E2).12,27 However, with an activation energy of only 0.8 eV below the conduction band, Fe has shown to be a source of device instabilities. For example, Mcglone et al. showed a clear link between the thermal emission of electrons from the EC-0.8 eV Fe state during device biasing and significant threshold voltage instabilities for β-Ga2O3 delta doped MESFETs.20 This results from Fe incorporation from the Fe-doped substrates tailing into the unintentionally doped (UID) epitaxial buffer layer between the substrate and the 2DEG device channel. Hence, while this degradation mode can be mitigated by using thicker UID buffers,19 this highlights the concern that the EC-0.8 eV position of Fe acceptors in the 4.6–4.8 eV bandgap of β-Ga2O3 may not be ideal since its charge state can be modulated during device biasing and trapped electrons can be thermally emitted. An impurity with a deeper acceptor level could therefore be advantageous in reducing the chance of modulating its charge state during biasing and becoming thermally ionized, even at the high operating temperatures envisioned for β-Ga2O3 power electronics.

As a result, there is interest in alternative acceptor-like dopants to Fe. A recent study published by Seyidov et al. demonstrated cobalt doping in Czochralski (CZ) growth, resulting in semi-insulating β-Ga2O3 substrates with the Fermi level pinned at EC-2.1 eV.13 Similarly, Mg doping has resulted in high-quality semi-insulating β-Ga2O3 materials with a predicted acceptor level at EC-3.4 eV by DFT.11,14,28 Nitrogen has been identified as a deep acceptor candidate by Peelaers et al. through DFT calculations.11 N substituting on oxygen I, II, and III sites in the β-Ga2O3 lattice forms acceptor-like levels with (0/−1) charge-state transition levels at EC-1.61 eV, EC-1.46 eV, and EC-2.84 eV, respectively.11 In recent experimental measurements after ion implantation, N was found to diffuse only at temperatures greater than 1100 °C, while Mg diffusion was observed at temperatures as low as 800 °C.29,30 Using these results and evaluating them with transition state theory, a barrier diffusion activation energy of 3.87 eV for N and 2.84 eV for Mg was obtained.11,29 Large activation energies indicate high thermal stability in extreme environments and are desirable for deep acceptor dopants. In addition, the ion implantation study reported that a high acceptor doping efficiency was achieved using N.29 Nitrogen has been successfully incorporated in epitaxial layers grown by metal-organic chemical vapor deposition (MOCVD) using an N2O precursor, resulting in a room-temperature electron mobility of 153 cm2/Vs and a low net ionized doping concentration of 1014 cm−3.26,31 With these promising preliminary work as motivation and given the lack of experimental knowledge regarding the nature of nitrogen-related defects within β-Ga2O3, we have undertaken a study to investigate the electronic properties associated with nitrogen-related deep-level defects within β-Ga2O3. A combination of both deep-level transient (thermal) spectroscopy (DLTS) and deep-level optical spectroscopy (DLOS) measurements was applied to a systematic sample set of Si and N co-doped β-Ga2O3 materials designed to allow for precise and quantitative identification of nitrogen-related defect states in β-Ga2O3, which were confirmed by DFT calculations. This study suggests very efficient incorporation of nitrogen into a preferred configuration in the β-Ga2O3 lattice, having an energy level of EC-2.9 eV, implying significant promise for future applications of nitrogen doping to create ideal, highly resistive β-Ga2O3 materials for implementation into future device technologies.

A systematic set of experiments was performed using several N-implant conditions on thick, Si-doped hydride vapor phase epitaxy (HVPE) grown β-Ga2O3 materials with (001) orientation to investigate the introduction of deep levels associated with nitrogen. The reference sample was a 10 μm thick Si-doped HVPE grown epilayer on a highly conductive (n = 3–5 × 1018 cm−3) Sn-doped β-Ga2O3 substrate procured from Novel Crystal Technology.32,33 The HVPE-grown epilayer was doped with Si to achieve a net ionized doping concentration of 1.2 × 1017 cm−3. This value was chosen so that subsequent nitrogen co-doping via ion implantation could be used to achieve nitrogen concentrations high enough to be measured using secondary ion mass spectrometry (SIMS), i.e., above the SIMS detection limit of ∼1 × 1016 cm−3, while simultaneously producing an appropriate degree of carrier compensation to be within ideal measurement conditions for capacitance-based DLTS and DLOS characterization. Using Stopping and Range of Ions in Matter (SRIM) simulations,34 a uniform implantation profile was created by combining nitrogen implant energies from 10 to 680 keV, each with different dose; an example of the individual implant profiles is shown in Fig. 1(a) in the light gray curves. The cumulative nitrogen implant profile from the individual energy implants results in the desired uniform nitrogen concentration profiles. Two sets of samples were created, one with a uniform nitrogen doping of 2 × 1016 and another at 5 × 1016 cm−3, within the top ∼0.8 µm of the HVPE layer, depicted in Fig. 1(a). The implantation was performed using a commercial implanter (Cutting Edge Ions). Post-implantation annealing was performed at 1100 °C in nitrogen ambient for 10 min using a rapid thermal annealing (RTA) system to remove implantation damage and activate the dopants as reported elsewhere.29 The annealing was performed on a ULVAC technology MILA 3000 rapid thermal annealing (RTA) system.

FIG. 1.

(a) Nitrogen implantation profiles simulated using SRIM and implemented by implanting with multiple energies into the Si-doped HVPE layer for the two target concentrations. The figure provides implant profiles of N for the individual implant energies from 10 to 680 keV used to achieve the uniform implantation profiles. Cross-sectional diagrams of the fabricated diode structures with Ni-Schottky contacts and backside ohmic on (b) the baseline Si-doped HVPE sample and (c) the N co-doped sample with the implanted region in the upper 0.8 μm are shown.

FIG. 1.

(a) Nitrogen implantation profiles simulated using SRIM and implemented by implanting with multiple energies into the Si-doped HVPE layer for the two target concentrations. The figure provides implant profiles of N for the individual implant energies from 10 to 680 keV used to achieve the uniform implantation profiles. Cross-sectional diagrams of the fabricated diode structures with Ni-Schottky contacts and backside ohmic on (b) the baseline Si-doped HVPE sample and (c) the N co-doped sample with the implanted region in the upper 0.8 μm are shown.

Close modal

Multiple Schottky diodes were fabricated on each of the three samples: a baseline HVPE sample without implantation and annealing and the two implanted and annealed samples with nitrogen concentrations of 2 × 1016 and 5 × 1016 cm−3 confirmed by SIMS (discussed later). The Schottky diode fabrication process in all cases is now described. All samples were solvent cleaned with acetone and isopropyl alcohol (IPA), followed by a rinse in deionized (DI) water. Semi-transparent 8 nm Ni Schottky contacts (290 × 290 μm2) were then deposited using electron (e)-beam evaporation, followed by an e-beam deposited ohmic contact stack of Ti/Al/Ni/Au (20/200/30/200 nm) on the back side of the conducting substrate. A thick Ni/Au (30/200 nm) contact stack of 25 × 25 μm2 was deposited on a small area of the thin 8 nm semi-transparent Ni-Schottky for probing the device and making stable contacts during measurements. A cross-sectional diagram depicting the fabricated diodes on the baseline HVPE and the post-implant annealed samples are shown in Figs. 1(b) and 1(c), respectively. The fabricated samples yielded dozens of high-quality Schottky diodes across each sample, with nearly identical I–V characteristics having ideality factors of 1.1 ± 0.05 and leakage current densities below 10−9 A/cm2. Typical Schottky barrier heights were found to be ∼1.4 V as measured using internal photoemission (IPE) spectroscopy, with negligible changes after N-implantation.35,36

DLTS measurements on the fabricated diodes were carried out over a temperature range from 77 to 400 K, enabling the determination of concentrations and activation energies for trap states within ∼1 eV of the conduction band. Briefly, the DLTS conditions used here consisted of a “fill pulse” of 0 V for 10 ms to fill traps, followed by a “quiescent state” reverse bias at −2 V for 2 s, during which any capacitance transients due to thermally stimulated electron emission from traps are recorded at each temperature from 77 to 400 K in steps of 0.1 K and subsequently analyzed using a conventional double boxcar method. The full details of the measurement can be found in previous publications.37–40 The remaining distribution of trap states in the bandgap was obtained using DLOS measurements. DLOS is based on direct photoemission of trapped carriers from states too deep to be seen by thermally stimulated carrier emission measurements like DLTS. The trap-filling in the depletion region is achieved with a fill pulse voltage of 0.7 V for 10 s, followed by a thermal settling time of 100 s. Trapped carriers, if any, are then optically emitted by a monochromatized light source providing photons with energies between 1.2 and 5.0 eV, scanning the bandgap in 0.02 eV steps for 300 s each, and photocapacitance transients are recorded. These are subsequently analyzed to extract optical cross sections of any traps present as a function of photon energy. The extracted optical cross sections are then fitted to the Pässler model, which enables the determination of the equilibrium trap energy level and the Frank–Condon energy (DFC).41,42 The concentration of each DLOS-detected state is determined from the steady-state photocapacitance (SSPC) associated with each state. The lighted C–V (LCV) method was also used to assure complete saturation of the SSPC for accurate trap concentration extraction since we have reported that for cases where a defect state possesses a small optical cross section, SSPC saturation can be difficult.35,36,43,44

As noted above, the Si:N co-doping concentrations were chosen to allow correlations between the nitrogen concentration, carrier compensation, and deep-level concentration by being within the measurement limits of both SIMS and DLTS/DLOS. The nitrogen concentration profiles obtained by SIMS for both target concentrations after the post-implant anneals are shown in Fig. 2(a), revealing uniform implant profiles in close agreement with the targeted design discussed earlier. Capacitance–voltage (C–V) measurements reveal systematically increased carrier compensation (lower net doping) as a function of increased nitrogen doping, as shown in Fig. 2(b) (it should be noted that without post-implant annealing, the implanted samples are fully depleted, presumably due to implantation damage). Based on the simple assumption that each nitrogen defect is activated and removes a single conduction band electron, nitrogen doping should remove ∼17% and 42% of the free electron concentration in the lower and higher nitrogen co-doped samples, respectively, for each N concentration. However, the results indicate that while almost all the implant damage was removed by the post-implant anneal, the slightly higher amount of carrier removal actually observed at 27% and 67%, respectively, suggests that some damage in the form of additional compensating defects is likely to remain, in addition to the nitrogen acceptors. This work, however, is not focused on anneal optimization. The next section presents comprehensive DLTS and DLOS investigations, which focus on identifying nitrogen-related states but can also reveal residual defect states that could contribute to the slightly higher degree of observed carrier compensation for the implant and anneal conditions used here.

FIG. 2.

(a) The N concentration profiles measured by SIMS in each sample and the (b) corresponding net ionized doping (ND–NA) concentrations extracted from the CV for each co-doped sample, after both implant and anneal, along with the non-implanted control (black) sample. The N concentration in (a) matches the difference in net ionized doping (b) between the control sample and the respective implanted samples in a monotonic fashion, indicating very efficient carrier removal by N, as desired.

FIG. 2.

(a) The N concentration profiles measured by SIMS in each sample and the (b) corresponding net ionized doping (ND–NA) concentrations extracted from the CV for each co-doped sample, after both implant and anneal, along with the non-implanted control (black) sample. The N concentration in (a) matches the difference in net ionized doping (b) between the control sample and the respective implanted samples in a monotonic fashion, indicating very efficient carrier removal by N, as desired.

Close modal

The DLTS spectra for the baseline and two N co-doped (post-implant and annealed) HVPE samples are shown in Figs. 3(a)3(c), with the Arrhenius characteristic for each DLTS trap shown in Fig. 3(d). The DLTS spectra reveal states at EC-0.25 eV, EC-0.34 eV, EC-0.4 eV, EC-0.6 eV, EC-0.7 eV, and EC-0.8 eV. The levels at EC-0.7 eV and EC-0.8 eV are also present in the baseline Si-doped HVPE sample, while the other four levels were only observed in the N co-doped samples. The trap concentrations for each state were calculated using the “lambda” correction to account for the volume of the depletion region in which the occupancy of the specific trap level is modulated by the DLTS biasing.39 This is critical for the most accurate concentration values, shown for each trap in Figs. 3(a)3(c).

FIG. 3.

DLTS spectra measured at a rate window of 0.8 s−1 on the (a) baseline HVPE sample, (b) 2 × 1016 cm−3 N co-doped sample, and (c) 5 × 1016 cm−3 N co-doped sample, with lambda corrected concentrations for each state presented in the trap labels. The Arrhenius plot in (d) compares the traps from (a)–(c) with previously published results of traps due to high energy proton irradiation damage and from Fe impurities in β-Ga2O3.7,12,21

FIG. 3.

DLTS spectra measured at a rate window of 0.8 s−1 on the (a) baseline HVPE sample, (b) 2 × 1016 cm−3 N co-doped sample, and (c) 5 × 1016 cm−3 N co-doped sample, with lambda corrected concentrations for each state presented in the trap labels. The Arrhenius plot in (d) compares the traps from (a)–(c) with previously published results of traps due to high energy proton irradiation damage and from Fe impurities in β-Ga2O3.7,12,21

Close modal

Several observations are important from the DLTS data. First, none of these energy levels match anywhere close to the predicted nitrogen states previously reported for β-Ga2O3. Hence, the sources of these states are likely due to pre-existing defects in the baseline Si-doped HVPE material or result from implantation damage that was not fully removed by the post-implant anneal process used here and is not related to nitrogen.7,12,21,45 Second, regardless of their source, the individual and total concentrations of DLTS-detected traps are on the order of 1013–1014 cm−3. This is far lower than what can account for the large carrier removal observed in the C–V data. However, the total is large enough to account for some of the differences between the anticipated and measured amount of carrier removal discussed above. While this work is not focused on optimizing the implant and annealing conditions for nitrogen doping by implantation, it is still helpful to consider possible sources for these defect states to provide insights into the HVPE material and the implant process used here.

To that end, we have included other sets of well-known and commonly observed DLTS data with the Arrhenius plot shown in Fig. 3(d), which are associated with Fe impurities and point defects due to high-energy particle irradiation, as labeled in the same figure. Comparisons with the data here are revealing. First, the radiation-induced defect data closely matches three trap states observed here at EC-0.34, EC-0.6, and EC-0.7 eV.6,7 Since the proton radiation damage produces intrinsic point defects, these are all likely due to residual damage from the nitrogen implantation process. The state at EC-0.7 eV has received significant focus in prior work and has been connected with gallium vacancies; recent work has proposed the Ga–O divacancy as the most likely origin.6,7,12,17 Additionally, the state detected at the EC-0.8 eV closely matches the well-known trap associated with FeGa defects, suggesting that residual Fe is present in the HVPE material.12,15,21 As noted above, both the EC-0.7 and EC-0.8 eV traps are present in the control baseline HVPE sample, and the concentration of the EC-0.8 eV trap did not change, consistent with its source being residual Fe. Finally, the sources of the remaining two DLTS states at EC-0.25 eV and EC-0.4 eV could be introduced by the implant damage and/or by the annealing process. Regardless, none of the DLTS traps have high enough concentrations of consequence to explain the strong carrier compensation observed in the co-doped samples. Thus, the section titled "DLOS results" focuses on the rest of the bandgap using DLOS measurements.

The individual DLTS measured trap concentrations (∼1013–1014 cm−3) were several orders of magnitude lower than the targeted N concentrations shown in Fig. 1(a) or the measured carrier compensation in Fig. 2(b). Furthermore, the N on oxygen substitutional sites in β-Ga2O3 form acceptor levels predicted by DFT to be significantly deeper than anything observed by DLTS. Therefore, DLOS and LCV measurements were used to interrogate the rest of the bandgap to investigate the possible presence of other state(s) responsible for the carrier compensation. The SSPC data in Fig. 4(a) reveal three SSPC onsets at EC-2.0 eV, EC-2.9 eV, and EC-4.4 eV, associated with very deep levels in the bandgap. The concentrations from the SSPC onsets indicate trap concentrations on the order of 1016 cm−3, far higher than the DLTS trap concentrations. In order to ensure complete saturation of the DLOS-detected states to compare concentrations with SIMS, LCV measurements were applied to each state. Considering the EC-2.9 eV state first, Fig. 4(b) compares independently measured EC-2.9 eV trap concentrations with SIMS-measured N concentration as a function of the targeted N implant concentration. There is a clear correlation and quantitative match between the concentration of this trap and the total N concentration. The LCV measured trap concentrations for the EC-2.9 eV state were found to be 1.8 × 1016 cm−3 and 4.5 × 1016 cm−3 for the 2 × 1016 cm−3 and 5 × 1016 cm−3 N co-doped samples.

FIG. 4.

(a) Steady state photocapacitance spectra measured on the baseline HVPE and N co-doped samples with the precise trap energy and DFC calculated from optical cross section fitting to the Pässler model.42 A monotonic increase in the trap concentration of EC-2.9 eV state with N is evident from the SSPC spectra. (b) This figure shows the match between the EC-2.9 eV trap concentration measured by LCV, the total N content from SIMS, and the target N implant concentration. The dotted line represents a one-to-one correspondence between the target N concentration and the measured concentrations of N and EC-2.9 eV trap.

FIG. 4.

(a) Steady state photocapacitance spectra measured on the baseline HVPE and N co-doped samples with the precise trap energy and DFC calculated from optical cross section fitting to the Pässler model.42 A monotonic increase in the trap concentration of EC-2.9 eV state with N is evident from the SSPC spectra. (b) This figure shows the match between the EC-2.9 eV trap concentration measured by LCV, the total N content from SIMS, and the target N implant concentration. The dotted line represents a one-to-one correspondence between the target N concentration and the measured concentrations of N and EC-2.9 eV trap.

Close modal

The overall quantitative trends for all detected states as a function of N implant concentrations are depicted in Fig. 5, confirming that the primary influence of N doping is on the EC-2.9 eV state. Before considering the physical defect configuration associated with the EC-2.9 eV state, the two DLOS detected states at EC-2.0 eV and EC-4.4 eV warrant some discussion since both display small but measurable increases in their concentrations with nitrogen implantation, as seen in Fig. 5. Previous work on gallium oxide subjected to high energy particle irradiation by both protons and neutrons has connected the EC-2.0 eV state with a 2VGa1Gai defect “split-vacancy” complex,46 and gallium vacancies will likely form due to the implantation damage (though largely annealed out). The commonly reported EC-4.4. eV state is not well understood, and its physical source is still being explored, with prior reports showing inconsistent dependencies on variables such as growth method and high energy particle irradiation conditions.6,7,36,44,47,48

FIG. 5.

Summary of the energy positions and concentrations for traps detected by DLTS and DLOS in baseline HVPE, 2 × 1016 cm−3, and 5 × 1016 cm−3 N co-doped HVPE samples. The lambda-corrected DLTS-detected trap concentrations are in the upper 1 eV from the conduction band, while the DLOS-detected trap concentrations are measured from LCV. EC-2.9 eV trap concentration measured from LCV tracks with an increase in nitrogen implant concentration highlighted in red color.

FIG. 5.

Summary of the energy positions and concentrations for traps detected by DLTS and DLOS in baseline HVPE, 2 × 1016 cm−3, and 5 × 1016 cm−3 N co-doped HVPE samples. The lambda-corrected DLTS-detected trap concentrations are in the upper 1 eV from the conduction band, while the DLOS-detected trap concentrations are measured from LCV. EC-2.9 eV trap concentration measured from LCV tracks with an increase in nitrogen implant concentration highlighted in red color.

Close modal

The above experimental studies strongly indicate that the EC-2.9 eV state is due to a nitrogen acceptor. To provide further verification and confirm the charge state and the possible physical defect configuration causing the emission process, we performed DFT studies. The calculations used projector augmented wave potentials49,50 as implemented in the Vienna Ab initio Simulation Package (VASP) code.51 Ga 3d electrons were treated as valence states. We used the Heyd–Scuseria–Ernzerhof (HSE) hybrid functional,52 with the screening parameter fixed at the standard value and the fraction of screened Hartree–Fock exchange set to 0.33, resulting in a direct bandgap of 4.9 eV. For nitrogen impurity calculations, we employed 160-atom supercells, a plane-wave energy cutoff of 500 eV, and a single special k-point at (0.25, 0.25, 0.25). Energies of various configurations and charge states were calculated using the formalism defined by Freysoldt et al.,53 including charge-state corrections.54–56 

As noted in the work of Peelaers et al.,11 the NO(III) configuration has the lowest energy among the three possible substitutional oxygen sites when the Fermi level is above 3 eV, a condition that is fulfilled in these Si-doped samples. As noted in the Introduction, the (0/−) charge-state transition levels for the nitrogen impurity were found to strongly depend on the site on which the nitrogen atom substitutes; again, it is the level for NO(III), at EC-2.84 eV, that closely matches the trap energy extracted from optical cross section shown in Fig. 6(a). In order to further confirm this identification, we calculated configuration coordinate diagrams57 in the one-dimensional approximation, which works very well when electron–phonon coupling is very strong as is clearly the case here. The diagram in Fig. 6(b) illustrates the potential energy surfaces as a function of atomic coordinates for the NO(III) impurity in the (ground-state, g) negative and (excited-state, e) neutral charge state, showing a large change in configuration coordinate ΔQ. The zero-phonon line (ZPL) energy EZPL corresponds to the charge-state transition level with respect to the conduction-band minimum, and effective phonon frequencies for ground and excited states are indicated. Within this model, the Franck–Condon (DFC) energy for the excitation process is the difference between the peak absorption energy Eabs and EZPL and thus given by 4.15–2.84 = 1.31 eV, in very good agreement with the DFC value extracted from the Pässler model shown in Fig. 6(a). 

FIG. 6.

(a) The optical cross section extracted from the photocapacitance transients of the N co-doped samples fitted to a Pässler model42 for calculating the equilibrium trap energy and DFC. EC-2.9 eV shows similar N-concentration as implanted and has a large DFC of 1.4 eV, implying broad lattice relaxation. (b) Configuration coordinate diagram calculated by DFT for NO(III) illustrating the optical transition that matches the optical cross section fitting results.

FIG. 6.

(a) The optical cross section extracted from the photocapacitance transients of the N co-doped samples fitted to a Pässler model42 for calculating the equilibrium trap energy and DFC. EC-2.9 eV shows similar N-concentration as implanted and has a large DFC of 1.4 eV, implying broad lattice relaxation. (b) Configuration coordinate diagram calculated by DFT for NO(III) illustrating the optical transition that matches the optical cross section fitting results.

Close modal

In conclusion, this study investigates the use of N as a deep acceptor dopant in β-Ga2O3 through ion implantation. Strong evidence of background carrier compensation with implanted N is presented, demonstrating over 95% dopant activation under optimized conditions. This work establishes a clear correlation between the nitrogen implant concentration and the EC-2.9 eV level determined by DLOS. As the nitrogen implantation concentration increases, there is a monotonic increase in the concentration of the EC-2.9 eV level, resulting in effective carrier compensation. First-principles DFT calculations indicate a (0/−1) transition for this state, verifying its acceptor-like behavior and showing that the experimentally observed configuration is for nitrogen occupying O site III (NO(III)), resulting in a deep compensating center at EC-2.9 eV. Furthermore, the photoionization of the NO(III) acceptor is accompanied by a large local lattice relaxation, leading to a high Franck–Condon (DFC) energy of 1.4 eV in the configuration coordinate diagram. This energy was experimentally obtained by analyzing the optical cross section derived from DLOS. This combined experimental and theoretical effort shows that nitrogen doping is a very effective process to introduce extremely efficient deep acceptors for the purpose of achieving very high degrees of controlled carrier compensation in β-Ga2O3.

The authors acknowledge the funding support from the Air Force Office of Scientific Research (Grant No. FA9550-18-1-0479) (Ali Sayir, Program Manager) and from the U.S. Air Force Radiation Effects Center of Excellence (Grant No. FA9550-22-1-0012). This work was supported, in part, by The Ohio State University Institute for Materials Research. This material is based on research sponsored by the Air Force Research Laboratory and Strategic Council for Higher Education under Agreement No. FA8650-19-2-9300 and the Ohio Department of Higher Education and Strategic Council for Higher Education under Ohio House Bill 49 of the 132nd General Assembly. The authors acknowledge the funding support from the Department of Energy/National Nuclear Security Administration under Award No. DE-NA0003921. This work was supported by the Air Force Office of Scientific Research (Grant No. FA9550-22-1-0165), the Research Council of Norway through the GO-POW project (Grant No. 314017), the University of Kansas General Research Fund allocation (Grant No. 2151089), and the Office of Naval Research/NRL Basic Research Program and partially performed under the auspices of the U.S. Department of Energy by Lawrence Livermore National Laboratory (LLNL) under Contract No. DE-AC52-07NA27344. The computations were performed on resources provided by UNINETT Sigma2, the National Infrastructure for High Performance Computing and Data Storage in Norway, and by the Extreme Science and Engineering Discovery Environment (XSEDE), supported by the National Science Foundation (Grant No. ACI-1548562). This work was supported by Air Force Office of Scientific Research Grant No. FA9550-18-1-0479 and National Science Foundation Award Nos. ECCS 2019749 and ECCS 2231026. Any opinions, findings, conclusions, or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the funding provider.

The authors have no conflicts to disclose.

Hemant Ghadi: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Investigation (lead); Methodology (lead); Validation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (lead). Joe F. McGlone: Formal analysis (supporting); Investigation (supporting); Writing – original draft (equal); Writing – review & editing (equal). Evan Cornuelle: Data curation (supporting); Formal analysis (supporting); Investigation (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Alexander Senckowski: Conceptualization (supporting); Software (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Shivam Sharma: Investigation (supporting); Resources (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Man Hoi Wong: Conceptualization (equal); Formal analysis (supporting); Investigation (supporting); Methodology (supporting); Resources (supporting); Supervision (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Uttam Singisetti: Investigation (supporting); Methodology (supporting); Resources (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Ymir Kalmann Frodason: Methodology (equal); Software (equal); Validation (equal); Writing – original draft (supporting); Writing – review & editing (supporting). Hartwin Peelaers: Methodology (supporting); Software (supporting); Validation (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). John L. Lyons: Methodology (supporting); Software (supporting); Validation (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Joel B. Varley: Methodology (equal); Software (equal); Validation (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Chris G. Van de Walle: Methodology (equal); Software (lead); Validation (equal); Writing – original draft (supporting); Writing – review & editing (supporting). Aaron Arehart: Investigation (supporting); Resources (supporting); Supervision (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Steven A. Ringel: Conceptualization (equal); Funding acquisition (lead); Investigation (equal); Project administration (lead); Resources (lead); Supervision (lead); Validation (supporting); Visualization (supporting); Writing – original draft (equal); Writing – review & editing (equal).

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

1.
H. H.
Tippins
, “
Optical absorption and photoconductivity in the band edge of β–Ga2O3
,”
Phys. Rev.
140
(
1A
),
A316
A319
(
1965
).
2.
M.
Orita
,
H.
Ohta
,
M.
Hirano
, and
H.
Hosono
, “
Deep-ultraviolet transparent conductive β-Ga2O3 thin films
,”
Appl. Phys. Lett.
77
(
25
),
4166
4168
(
2000
).
3.
T.
Onuma
,
S.
Saito
,
K.
Sasaki
,
T.
Masui
,
T.
Yamaguchi
,
T.
Honda
, and
M.
Higashiwaki
, “
Valence band ordering in β-Ga2O3 studied by polarized transmittance and reflectance spectroscopy
,”
Jpn. J. Appl. Phys.
54
(
11
),
112601
(
2015
).
4.
M. J.
Tadjer
,
J. L.
Lyons
,
N.
Nepal
,
J. A.
Freitas
,
A. D.
Koehler
, and
G. M.
Foster
, “
Editors’ choice–review–theory and characterization of doping and defects in β-Ga2O3
,”
ECS J. Solid State Sci. Technol.
8
(
7
),
Q3187
Q3194
(
2019
).
5.
S. J.
Pearton
,
J.
Yang
,
P. H.
Cary
,
F.
Ren
,
J.
Kim
,
M. J.
Tadjer
, and
M. A.
Mastro
, “
A review of Ga2O3 materials, processing, and devices
,”
Appl. Phys. Rev.
5
(
1
),
011301
(
2018
).
6.
H. J.
Ghadi
,
J. F.
McGlone
,
E.
Farzana
,
A. R.
Arehart
, and
S. A.
Ringel
, “
Radiation effects on β-Ga2O3 materials and devices
,” in
Ultrawide Bandgap β-Ga2O3 Semiconductor: Theory and Applications
, edited by
J. S.
Speck
and
E.
Farzana
(
AIP Publishing LLC
,
2023
).
7.
J. F.
McGlone
,
H.
Ghadi
,
E.
Cornuelle
,
A.
Armstrong
,
G.
Burns
,
Z.
Feng
,
A. F. M. A.
Uddin Bhuiyan
,
H.
Zhao
,
A. R.
Arehart
, and
S. A.
Ringel
, “
Proton radiation effects on electronic defect states in MOCVD-grown (010) β-Ga2O3
,”
J. Appl. Phys.
133
(
4
),
045702
(
2023
).
8.
Z.
Xia
,
H.
Xue
,
C.
Joishi
,
J.
Mcglone
,
N. K.
Kalarickal
,
S. H.
Sohel
,
M.
Brenner
,
A.
Arehart
,
S.
Ringel
,
S.
Lodha
,
W.
Lu
, and
S.
Rajan
, “
β-Ga2O3 delta-doped field-effect transistors with current gain cutoff frequency of 27 GHz
,”
IEEE Electron Device Lett.
40
(
7
),
1052
1055
(
2019
).
9.
Y.
Zhang
,
Z.
Xia
,
C.
Joishi
, and
S.
Rajan
, in
2018 76th Device Research Conference DRC
(
Material Research Society
,
2018
), pp.
1
2
.
10.
K. D.
Chabak
,
J. P.
McCandless
,
N. A.
Moser
,
A. J.
Green
,
K.
Mahalingam
,
A.
Crespo
,
N.
Hendricks
,
B. M.
Howe
,
S. E.
Tetlak
,
K.
Leedy
,
R. C.
Fitch
,
D.
Wakimoto
,
K.
Sasaki
,
A.
Kuramata
, and
G. H.
Jessen
, “
Recessed-gate enhancement-mode β-Ga2O3 MOSFETs
,”
IEEE Electron Device Lett.
39
(
1
),
67
70
(
2018
).
11.
H.
Peelaers
,
J. L.
Lyons
,
J. B.
Varley
, and
C. G.
Van de Walle
, “
Deep acceptors and their diffusion in Ga2O3
,”
APL Mater.
7
(
2
),
022519
(
2019
).
12.
M. E.
Ingebrigtsen
,
J. B.
Varley
,
A. Y.
Kuznetsov
,
B. G.
Svensson
,
G.
Alfieri
,
A.
Mihaila
,
U.
Badstübner
, and
L.
Vines
, “
Iron and intrinsic deep level states in Ga2O3
,”
Appl. Phys. Lett.
112
(
4
),
042104
(
2018
).
13.
P.
Seyidov
,
J. B.
Varley
,
Z.
Galazka
,
T.-S.
Chou
,
A.
Popp
,
A.
Fiedler
, and
K.
Irmscher
, “
Cobalt as a promising dopant for producing semi-insulating β-Ga2O3 crystals: Charge state transition levels from experiment and theory
,”
APL Mater.
10
(
11
),
111109
(
2022
).
14.
J. L.
Lyons
, “
A survey of acceptor dopants for β-Ga2O3
,”
Semicond. Sci. Technol.
33
(
5
),
05LT02
(
2018
).
15.
K.
Irmscher
,
Z.
Galazka
,
M.
Pietsch
,
R.
Uecker
, and
R.
Fornari
, “
Electrical properties of β-Ga2O3 single crystals grown by the Czochralski method
,”
J. Appl. Phys.
110
(
6
),
063720
(
2011
).
16.
Z.
Galazka
,
R.
Uecker
,
K.
Irmscher
,
M.
Albrecht
,
D.
Klimm
,
M.
Pietsch
,
M.
Brützam
,
R.
Bertram
,
S.
Ganschow
, and
R.
Fornari
, “
Czochralski growth and characterization of β-Ga2O3 single crystals
,”
Cryst. Res. Technol.
45
(
12
),
1229
1236
(
2010
).
17.
N. T.
Son
,
Q. D.
Ho
,
K.
Goto
,
H.
Abe
,
T.
Ohshima
,
B.
Monemar
,
Y.
Kumagai
,
T.
Frauenheim
, and
P.
Deák
, “
Electron paramagnetic resonance and theoretical study of gallium vacancy in β-Ga2O3
,”
Appl. Phys. Lett.
117
(
3
),
032101
(
2020
).
18.
S.
Bhandari
and
M. E.
Zvanut
, “
Charge trapping at Fe due to midgap levels in Ga2O3
,”
J. Appl. Phys.
129
(
8
),
085703
(
2021
).
19.
J. F.
McGlone
,
Z.
Xia
,
C.
Joishi
,
S.
Lodha
,
S.
Rajan
,
S.
Ringel
, and
A. R.
Arehart
, “
Identification of critical buffer traps in Si δ-doped β-Ga2O3 MESFETs
,”
Appl. Phys. Lett.
115
(
15
),
153501
(
2019
).
20.
J. F.
Mcglone
,
Z.
Xia
,
Y.
Zhang
,
C.
Joishi
,
S.
Lodha
,
S.
Rajan
,
S. A.
Ringel
, and
A. R.
Arehart
, “
Trapping effects in Si δ-doped β-Ga2O3 MESFETs on an Fe-doped β-Ga2O3 substrate
,”
IEEE Electron Device Lett.
39
(
7
),
1042
1045
(
2018
).
21.
H.
Ghadi
,
J. F.
McGlone
,
E.
Cornuelle
,
Z.
Feng
,
Y.
Zhang
,
L.
Meng
,
H.
Zhao
,
A. R.
Arehart
, and
S. A.
Ringel
, “
Deep level defects in low-pressure chemical vapor deposition grown (010) β-Ga2O3
,”
APL Mater.
10
(
10
),
101110
(
2022
).
22.
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
, “
Optical signatures of deep level defects in Ga2O3
,”
Appl. Phys. Lett.
112
(
24
),
242102
(
2018
).
23.
I.
Hany
,
G.
Yang
,
C. E.
Zhou
,
C.
Sun
,
K.
Gundogdu
,
D.
Seyitliyev
,
E. O.
Danilov
,
F. N.
Castellano
,
D.
Sun
, and
E.
Vetter
, “
Low temperature cathodoluminescence study of Fe-doped β-Ga2O3
,”
Mater. Lett.
257
,
126744
(
2019
).
24.
A.
Mauze
,
Y.
Zhang
,
T.
Mates
,
F.
Wu
, and
J. S.
Speck
, “
Investigation of unintentional Fe incorporation in (010) β-Ga2O3 films grown by plasma-assisted molecular beam epitaxy
,”
Appl. Phys. Lett.
115
(
5
),
052102
(
2019
).
25.
Z.
Wen
,
K.
Khan
,
X.
Zhai
, and
E.
Ahmadi
, “
Si doping of β-Ga2O3 by disilane via hybrid plasma-assisted molecular beam epitaxy
,”
Appl. Phys. Lett.
122
(
8
),
082101
(
2023
).
26.
F.
Alema
,
Y.
Zhang
,
A.
Osinsky
,
N.
Valente
,
A.
Mauze
,
T.
Itoh
, and
J. S.
Speck
, “
Low temperature electron mobility exceeding 104 cm2/V s in MOCVD grown β-Ga2O3
,”
APL Mater.
7
(
12
),
121110
(
2019
).
27.
A. Y.
Polyakov
,
N. B.
Smirnov
,
I. V.
Shchemerov
,
S. J.
Pearton
,
F.
Ren
,
A. V.
Chernykh
, and
A. I.
Kochkova
, “
Electrical properties of bulk semi-insulating β-Ga2O3 (Fe)
,”
Appl. Phys. Lett.
113
(
14
),
142102
(
2018
).
28.
Q. D.
Ho
,
T.
Frauenheim
, and
P.
Deák
, “
Theoretical confirmation of the polaron model for the Mg acceptor in β-Ga2O3
,”
J. Appl. Phys.
124
(
14
),
145702
(
2018
).
29.
M. H.
Wong
,
C.-H.
Lin
,
A.
Kuramata
,
S.
Yamakoshi
,
H.
Murakami
,
Y.
Kumagai
, and
M.
Higashiwaki
, “
Acceptor doping of β-Ga2O3 by Mg and N ion implantations
,”
Appl. Phys. Lett.
113
(
10
),
102103
(
2018
).
30.
C.
De Santi
,
M.
Fregolent
,
M.
Buffolo
,
M. H.
Wong
,
M.
Higashiwaki
,
G.
Meneghesso
,
E.
Zanoni
, and
M.
Meneghini
, “
Carrier capture kinetics, deep levels, and isolation properties of β-Ga2O3 Schottky-barrier diodes damaged by nitrogen implantation
,”
Appl. Phys. Lett.
117
(
26
),
262108
(
2020
).
31.
F.
Alema
,
Y.
Zhang
,
A.
Osinsky
,
N.
Orishchin
,
N.
Valente
,
A.
Mauze
, and
J. S.
Speck
, “
Low 1014 cm−3 free carrier concentration in epitaxial β-Ga2O3 grown by MOCVD
,”
APL Mater.
8
(
2
),
021110
(
2020
).
32.
K.
Nomura
,
K.
Goto
,
R.
Togashi
,
H.
Murakami
,
Y.
Kumagai
,
A.
Kuramata
,
S.
Yamakoshi
, and
A.
Koukitu
, “
Thermodynamic study of β-Ga2O3 growth by halide vapor phase epitaxy
,”
J. Cryst. Growth
405
,
19
22
(
2014
).
33.
A.
Kuramata
,
K.
Koshi
,
S.
Watanabe
,
Y.
Yamaoka
,
T.
Masui
, and
S.
Yamakoshi
, “
High-quality β-Ga2O3 single crystals grown by edge-defined film-fed growth
,”
Jpn. J. Appl. Phys.
55
(
12
),
1202A2
(
2016
).
34.
J. F.
Ziegler
,
M. D.
Ziegler
, and
J. P.
Biersack
, “
SRIM—The stopping and range of ions in matter
,”
Nucl. Instrum. Methods Phys. Res., Sect. B
268
,
1818
1823
(
2010
).
35.
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
, “
Full bandgap defect state characterization of β-Ga2O3 grown by metal organic chemical vapor deposition
,”
APL Mater.
8
(
2
),
021111
(
2020
).
36.
H.
Ghadi
,
J. F.
McGlone
,
Z.
Feng
, and
S. A.
Ringel
, “
Influence of growth temperature on defect states throughout the bandgap of MOCVD-grown β-Ga2O3
,”
Appl. Phys. Lett.
117
,
172106
(
2020
).
37.
Z.
Zhang
,
E.
Farzana
,
A. R.
Arehart
, and
S. A.
Ringel
, “
Deep level defects throughout the bandgap of (010) β-Ga2O3 detected by optically and thermally stimulated defect spectroscopy
,”
Appl. Phys. Lett.
108
(
5
),
052105
(
2016
).
38.
H.
Ghadi
,
J. F.
McGlone
,
Z.
Feng
,
A. F. M. A. U.
Bhuiyan
,
Y.
Zhang
,
H.
Zhao
,
A.
Armstrong
,
G. R.
Burns
,
G.
Vizkelethy
,
E.
Bielejec
,
A. R.
Arehart
, and
S. A.
Ringel
, “
Oxide-based mater and devices XII
,”
Proc. SPIE
11687
,
1168701
(
2021
).
39.
P.
Blood
and
J. W.
Orton
,
The Electrical Characterization of Semiconductors: Majority Carriers and Electron States
(
Academic Press Limited
,
San Diego, CA
,
1992
).
40.
D. V.
Lang
, “
Deep-level transient spectroscopy: A new method to characterize traps in semiconductors
,”
J. Appl. Phys.
45
(
7
),
3023
3032
(
1974
).
41.
A.
Chantre
,
G.
Vincent
, and
D.
Bois
, “
Deep-level optical spectroscopy in GaAs
,”
Phys. Rev. B
23
(
10
),
5335
5359
(
1981
).
42.
R.
Pässler
, “
Photoionization cross-section analysis for a deep trap contributing to current collapse in GaN field-effect transistors
,”
J. Appl. Phys.
96
(
1
),
715
722
(
2004
).
43.
A.
Armstrong
,
A. R.
Arehart
, and
S. A.
Ringel
, “
A method to determine deep level profiles in highly compensated, wide band gap semiconductors
,”
J. Appl. Phys.
97
(
8
),
083529
(
2005
).
44.
A. M.
Armstrong
,
M. H.
Crawford
,
A.
Jayawardena
,
A.
Ahyi
, and
S.
Dhar
, “
Role of self-trapped holes in the photoconductive gain of β-gallium oxide Schottky diodes
,”
J. Appl. Phys.
119
,
103102
(
2016
).
45.
M. E.
Ingebrigtsen
,
A. Y.
Kuznetsov
,
B. G.
Svensson
,
G.
Alfieri
,
A.
Mihaila
,
U.
Badstübner
,
A.
Perron
,
L.
Vines
, and
J. B.
Varley
, “
Impact of proton irradiation on conductivity and deep level defects in β-Ga2O3
,”
APL Mater.
7
(
2
),
022510
(
2019
).
46.
J. M.
Johnson
,
Z.
Chen
,
J. B.
Varley
,
C. M.
Jackson
,
E.
Farzana
,
Z.
Zhang
,
A. R.
Arehart
,
H.-L.
Huang
,
A.
Genc
,
S. A.
Ringel
,
C. G.
Van de Walle
,
D. A.
Muller
, and
J.
Hwang
, “
Unusual formation of point-defect complexes in the ultra-wide band-gap semiconductor β-Ga2O3
,”
Phys. Rev. X
9
(
4
),
041027
(
2019
).
47.
E.
Farzana
,
A.
Mauze
,
J. B.
Varley
,
T. E.
Blue
,
J. S.
Speck
,
A. R.
Arehart
, and
S. A.
Ringel
, “
Influence of neutron irradiation on deep levels in Ge-doped (010) β-Ga2O3 layers grown by plasma-assisted molecular beam epitaxy
,”
APL Mater.
7
(
12
),
121102
(
2019
).
48.
E.
Farzana
,
M. F.
Chaiken
,
T. E.
Blue
,
A. R.
Arehart
, and
S. A.
Ringel
, “
Impact of deep level defects induced by high energy neutron radiation in β-Ga2O3
,”
APL Mater.
7
(
2
),
022502
(
2019
).
49.
P. E.
Blöchl
, “
Projector augmented-wave method
,”
Phys. Rev. B
50
(
24
),
17953
17979
(
1994
).
50.
G.
Kresse
and
D.
Joubert
, “
From ultrasoft pseudopotentials to the projector augmented-wave method
,”
Phys. Rev. B
59
(
3
),
1758
1775
(
1999
).
51.
G.
Kresse
and
J.
Furthmüller
, “
Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set
,”
Phys. Rev. B
54
(
16
),
11169
11186
(
1996
).
52.
A. V.
Krukau
,
O. A.
Vydrov
,
A. F.
Izmaylov
, and
G. E.
Scuseria
, “
Influence of the exchange screening parameter on the performance of screened hybrid functionals
,”
J. Chem. Phys.
125
(
22
),
224106
(
2006
).
53.
C.
Freysoldt
,
B.
Grabowski
,
T.
Hickel
,
J.
Neugebauer
,
G.
Kresse
,
A.
Janotti
, and
C. G.
Van de Walle
, “
First-principles calculations for point defects in solids
,”
Rev. Mod. Phys.
86
(
1
),
253
305
(
2014
).
54.
C.
Freysoldt
,
J.
Neugebauer
, and
C. G.
Van de Walle
, “
Fully ab initio finite-size corrections for charged-defect supercell calculations
,”
Phys. Rev. Lett.
102
(
1
),
016402
(
2009
).
55.
Y.
Kumagai
and
F.
Oba
, “
Electrostatics-based finite-size corrections for first-principles point defect calculations
,”
Phys. Rev. B
89
(
19
),
195205
(
2014
).
56.
T.
Gake
,
Y.
Kumagai
,
C.
Freysoldt
, and
F.
Oba
, “
Finite-size corrections for defect-involving vertical transitions in supercell calculations
,”
Phys. Rev. B
101
(
2
),
020102
(
2020
).
57.
A.
Alkauskas
,
J. L.
Lyons
,
D.
Steiauf
, and
C. G.
Van de Walle
, “
First-principles calculations of luminescence spectrum line shapes for defects in semiconductors: The example of GaN and ZnO
,”
Phys. Rev. Lett.
109
(
26
),
267401
(
2012
).