Epitaxial layers of α-Ga2O3 with different Sn doping levels were grown by halide vapor phase epitaxy on sapphire. The films had shallow donor concentrations ranging from 1017 to 4.8 × 1019 cm−3. Deep level transient spectroscopy of the lowest doped samples revealed dominant A traps with level Ec − 0.6 eV and B traps near Ec − 1.1 eV. With increasing shallow donor concentration, the density of the A traps increased, and new traps C (Ec − 0.85 eV) and D (Ec − 0.23 eV) emerged. Photocapacitance spectra showed the presence of deep traps with optical ionization energy of ∼2 and 2.7 eV and prominent persistent photocapacitance at low temperature, surviving heating to temperatures above room temperature. The diffusion length of nonequilibrium charge carriers was 0.15 µm, and microcathodoluminescence spectra showed peaks in the range 339–540 nm, but no band-edge emission.

Ga2O3 and related ternary solid solutions are promising next-generation materials for use in high-power devices. This is due to their excellent properties, including band gaps close to 5 eV, electric breakdown field about 3 times higher than for SiC and GaN, and electron saturation velocity similar to SiC and GaN.1–4 The main focus in Ga2O3 materials and device research has been on the thermodynamically stable, monoclinic β-polytype, for which large diameter, high crystalline quality bulk crystals, epitaxial films, and heterojunctions exist.1–5 This has led to demonstrations of prototype power rectifiers, field effect transistors, and solar-blind photodetectors.2,4,5

However, other Ga2O3 polytypes are also of interest. The most important among those seems to be the α-polytype with corundum crystalline structure.1 The band gap of α-Ga2O3 is higher than β-Ga2O3, about 5.2 eV vs 4.6–4.8 eV.1,2 The higher symmetry of the rhombohedral corundum structure is an additional bonus compared to the monoclinic structure of β-Ga2O3, where the anisotropy of many properties is well documented.6–8 Although α-Ga2O3 is not thermodynamically stable, its formation energy is only slightly higher than for β-Ga2O3 and formation of the single crystalline α-polytype can be achieved under nonequilibrium growth conditions. An additional attractive feature is the compatibility of α-Ga2O3 growth with deposition on other corundum-structured materials, most importantly sapphire, which is cheap, readily available in large diameters, and with excellent crystalline quality. A problem with all Ga2O3 polytypes is the lack of suitable p-type dopants in Ga2O3.2–5 The way around this proposed so far is the use of Ga2O3 heterojunctions with wide-band gap materials with controlled p-type conductivity.2 Although the formation of n-type α-Ga2O3 heterojunctions with p-type wide-band gap materials, such as copper oxide, nickel oxide, SiC, and GaN, have been reported,9 the processes involved are technologically challenging due to the difference in the lattice parameters, type of crystalline structure, and the necessity to overcome the formation of highly defective interfacial layers.9 High growth temperature of β-Ga2O3 also seems to present difficulties when combining gallium oxide with other wide-band gap materials allowing effective p-type doping.9 In that sense, the corundum structure is more compatible with growth on these materials, a better matching of in-plane lattice constants of α-Ga2O3 with advanced hexagonal structure wide-band gap materials, such as SiC or GaN; a typically lower growth temperature of α-polytype can make high-quality heterojunction preparation easier. It also helps that related corundum-structured oxides, such as α-(Ga1−xIrx)2O3 or α-(Ga1−xRhx)2O3,10,11 have been reported to possess well-defined p-type conductivity. Moreover, because of the similarity in symmetry and the relative closeness of in-plane lattice parameters for the (00.1) plane of α-Ga2O3 and (111) plane of diamond, it could prove to be possible to grow high-quality heterojunctions between the n-type α-Ga2O3 and p-type or semi-insulating (111) diamond. If successful, this would help to overcome another serious drawback of Ga2O3, the relatively low thermal conductivity1–5 that is a disadvantage for power devices.

There is still the challenge of growing single polytype α-Ga2O3 films with high crystalline quality and controlled doping, given that growth is on alien substrates, most commonly sapphire. Recently, high crystalline quality single polytype α-Ga2O3 films were grown by mist chemical vapor deposition (MIST CVD), a version of metalorganic chemical vapor deposition (MOCVD), on basal plane sapphire.10,11 The crystalline quality can be further improved by using halide vapor phase epitaxy (HVPE) epitaxial lateral overgrowth (ELO) for masked growth on sapphire.12 Dislocation densities <5 × 106 cm−2 in the window regions of such structures have been reported, although growth of α-Ga2O3 was competing with growth of the hexagonal ε-polytype.12 Even without ELO, the crystalline quality of α-Ga2O3 films grown by HVPE on sapphire improves as the film thickness increases from ∼2 µm to 12 µm.13 

Pulsed HVPE growth of α-Ga2O3 on sapphire produces films with screw dislocation densities of (2–4) × 105 cm−2 and edge dislocations on the order of 109 cm−2.14 Doping with Sn donors in MIST CVD produced n-type material with shallow donor densities controlled from 1017–1019 cm−310,11 while undoped films prepared by either MIST CVD or by different versions of HVPE were highly resistive.14 These developments have made it possible to demonstrate α-Ga2O3-based power rectifiers with attractive performance: low on-resistance of 0.1–0.4 mΩ cm2 and breakdown voltage of 531–855 V,10,11 although the authors had to use lift-off and transfer of the grown device structures to overcome the heat control problems. However, little is known about the deep traps, diffusion lengths of nonequilibrium charge carriers, and luminescence spectra of doped and undoped α-Ga2O3 films, in contrast to abundant data for the β-polytype.23–26 In undoped semi-insulating (SI) α-Ga2O3 films, deep traps near Ec − 1 eV pinning the Fermi level, deep traps with levels near Ev + 1.4 eV, and shallower electron traps near Ec − 0.3 eV and Ec − 0.6 eV, were reported.14 In this article, we present deep trap and minority carrier transport studies for Sn-doped α-Ga2O3 films grown by HVPE on basal plane sapphire.

Ga2O3 films were grown on c-plane sapphire substrates in a hot wall atmospheric pressure HVPE reactor with a multizone horizontal furnace. GaCl and oxygen were used as precursors. Sn was used as the n-type dopant. The GaCl vapor was synthesized in situ upstream in the reactor by the reaction of metallic gallium (99.9999%) and gaseous hydrogen chloride (99.999%). The yield of GaCl formation is estimated as >80%. The GaCl vapor was transported to the deposition zone of the reactor held at 500 °C, where it reacted with O2 to produce Ga2O3. The HCl flow through the source was varied from 0.1 to 0.4 slm. The oxygen flow was constant at 0.2 slm. Argon was used as a carrier gas and the total gas flow through the reactor was up to 10 slm. Under these conditions, the deposition rate was 5–10 µm/h. The deposition time was adjusted to obtain Ga2O3 films with thickness 2–12 µm. Specifically, the Sn-doped films had thickness ∼2 µm. Three samples with n-type doping of about 1017 cm−3, 5 × 1017 cm−3, and 1019 cm−3 were prepared. After growth, the substrates were cooled to room temperature under a flow of Ar.

The samples were characterized by ω − 2Θ x-ray scans that showed all the films to be single-phase α-Ga2O3 with (00.1) orientation. Structural characterization included high resolution x-ray diffraction (HRXRD) rocking curves full width at half maximum (FWHM) for the (00.6) reflections,15 surface morphology observation in secondary electron (SE) mode of scanning electron microscope (SEM), microcathodoluminescence (MCL) at 90 and 300 K, electron beam induced current (EBIC) imaging of the surface with deposited Schottky diodes, and diffusion length measurements using the EBIC collection efficiency dependence on accelerating voltage of the probing SEM beam. Experimental setups were described previously.15–17 The model used to determine the diffusion length from EBIC has been developed for Ga2O316 and used for diffusion lengths in HVPE-grown β-Ga2O3.18,19

Electrical measurements were performed using semitransparent 1-mm diameter Ni Schottky diodes and Ti/Au Ohmic contacts.14 These measurements involved capacitance-voltage (C-V) measurements at 20 Hz–1 MHz in the dark and under monochromatic illumination with high-power light emitting diodes (LEDs) with wavelengths from 365 to 940 nm, capacitance vs frequency (C-f), admittance spectra (AS),20 deep level spectra with electrical (DLTS), or optical (ODLTS) injection.21–24 

All samples had thickness ∼2 µm, with visually smooth surfaces. The results of HRXRD FWHM of the (00.6) HRXRD rocking curves measurements were close to 0.35°–0.43° (Table I), indicating a dislocation density >109 cm−2. For the most lightly Sn-doped sample, sample 1, EBIC imaging showed large dark defects, also seen in SE images. These are most likely growth pores. In addition, there are a large number of small dark defects, forming a cellular structure likely due to dislocation grain boundaries [Fig. 1(top)]. The density of these defects is slightly higher and the number of pores much lower for the heavily Sn-doped sample [Fig. 1 (bottom)]. Thick (10–12 µm) undoped semi-insulating samples demonstrate better characteristics, with FWHM typically 0.12°,9 i.e., three times lower than for the present set.

TABLE I.

Characteristics of the α-Ga2O3 HVPE samples.

Deep traps concentrations (cm−3)
Sample Nos.(00.6) FWHM (deg)Nd (cm−3)A (Ec − 0.6 eV)B (Ec − 1.1 eV)C (Ec − 0.85 eV)D (Ec − 0.23 eV)
Sample 1 0.35 1.1 × 1017 5.3 × 1014 2.2 × 1015   
Sample 2 0.35 5 × 1017 2.3 × 1015  1.3 × 1015  
Sample 3 0.43 4.8 × 1019 5.8 × 1016   2 × 1016 
Deep traps concentrations (cm−3)
Sample Nos.(00.6) FWHM (deg)Nd (cm−3)A (Ec − 0.6 eV)B (Ec − 1.1 eV)C (Ec − 0.85 eV)D (Ec − 0.23 eV)
Sample 1 0.35 1.1 × 1017 5.3 × 1014 2.2 × 1015   
Sample 2 0.35 5 × 1017 2.3 × 1015  1.3 × 1015  
Sample 3 0.43 4.8 × 1019 5.8 × 1016   2 × 1016 
FIG. 1.

(Top) EBIC images of sample 1 (1017 cm−3); (bottom) sample 3 (4.8 × 1019 cm−3); the SEM beam current 1 nA, the accelerating voltage 30 kV.

FIG. 1.

(Top) EBIC images of sample 1 (1017 cm−3); (bottom) sample 3 (4.8 × 1019 cm−3); the SEM beam current 1 nA, the accelerating voltage 30 kV.

Close modal

All Sn-doped samples showed clear n-type conductivity from C-f and C-V results. The C-f characteristics showed a plateau at low frequencies and roll-off due to the effect of series resistance [Fig. 2(a)]. Current-voltage measurements of the three samples also indicated high series resistances, with the saturation current density Js = 7.1 × 10−7 A/cm2, ideality factor 2.1, reverse current density at −2 V of 4 × 10−4 A/cm2, series resistance Rs = 2.1 × 103 Ω for the least heavily doped sample, sample 1, and saturation current density 3.2 × 10−5 A/cm2, ideality factor 2.5, reverse current density at −2 V of 5.3 × 10−2 A/cm2, series resistance Rs = 210 Ω for most heavily doped sample, sample 3. This suggests that meaningful C-V and DLTS measurements have to be performed at frequencies below 10–20 kHz, which poses a problem with standard DLTS spectrometers but can be handled in our DLTS setup.22 For the two more heavily doped samples, the impact of leakage current on capacitance measurements at low frequencies limits the voltages to which the profiles can be measured and the temperatures to which the deep trap spectra can be reliably obtained. The 1/C2 vs voltage plots were linear when measured at frequencies corresponding to the C-f plateau [see Fig. 2(b) obtained for sample 1 at 10 kHz]. The net shallow donor concentration, Nd, calculated from these characteristics was 1.1 × 1017 cm−3 for sample 1, 5 × 1017 cm−3 for sample 2, and 4.8 × 1019 cm−3 for sample 3 (see Table I). The voltage offset of the 1/C2 vs V that should be close to the Schottky barrier height21 was 2.3 V at room temperature, slightly increased to 2.35 V at 80 K, and slightly decreased to 2.1 V at 400 K for the most lightly doped sample 1 [Fig. 2(b)]. The observed barrier height changes most likely reflect the change of the band gap with temperature.

FIG. 2.

(a) 300 K C-f dependences for sample 1 (1017 cm−3, red line), sample 2 (5 × 1017 cm−3, blue line), and sample 3 (4.8 × 1019 cm−3; olive line); (b) 1/C2 vs voltage plots for sample 1 obtained at 10 kHz for 294 K (black line), 84 K (blue line), and 402 K (red line).

FIG. 2.

(a) 300 K C-f dependences for sample 1 (1017 cm−3, red line), sample 2 (5 × 1017 cm−3, blue line), and sample 3 (4.8 × 1019 cm−3; olive line); (b) 1/C2 vs voltage plots for sample 1 obtained at 10 kHz for 294 K (black line), 84 K (blue line), and 402 K (red line).

Close modal

The two lower doped samples showed photocapacitance for photon energies higher than 2 eV (for the heavily doped sample 3, the effects were qualitatively similar but reliable quantitative measurements were difficult because of the low photocapacitance compared to the dark capacitance). At low temperatures, the photocapacitance became persistent, i.e., not returning to the dark value for a long time. Figure 3(a) shows 80 K concentration profiles measured in sample 1 in the dark and under illumination with photons of various energies produced by high-power GaN-based LEDs (output power ∼250 mW for all LEDs). Figure 3(b) shows the spectrum of photogenerated carriers calculated as the photoconcentration in the nearly flat region of the profiles [we took it for consistency at depth of 0.2 µm in Fig. 3(a)] from which the dark concentration is subtracted. The photocapacitance did not change even after keeping the samples in the dark for an hour.

FIG. 3.

(a) 80 K C-V profiles for sample 1 in the dark (black line), with 530 nm LED (olive line), 455 nm LED (cyan line), 400 nm LED (blue line), 380 nm LED (orange line), and 365 nm LED (violet line) illuminations; (b) spectra of concentration under illumination (red open squares), PPC concentration (open blue circles), and PPC + 2 V(open green triangles); (c) 80 K profiles in the dark (black line), with 400 nm LED illumination (solid blue line), after 400 nm LED illumination and 10 min in the dark (PPC curve, dashed blue line), and after additional application of forward bias pulse of 2 V (PPC + 2 V curve, dashed-dotted line, blue); (d) capacitance vs temperature dependence at 1 kHz for sample 2 in the dark (blue curve) and after illumination at 80 K with 365 nm LED (red curve).

FIG. 3.

(a) 80 K C-V profiles for sample 1 in the dark (black line), with 530 nm LED (olive line), 455 nm LED (cyan line), 400 nm LED (blue line), 380 nm LED (orange line), and 365 nm LED (violet line) illuminations; (b) spectra of concentration under illumination (red open squares), PPC concentration (open blue circles), and PPC + 2 V(open green triangles); (c) 80 K profiles in the dark (black line), with 400 nm LED illumination (solid blue line), after 400 nm LED illumination and 10 min in the dark (PPC curve, dashed blue line), and after additional application of forward bias pulse of 2 V (PPC + 2 V curve, dashed-dotted line, blue); (d) capacitance vs temperature dependence at 1 kHz for sample 2 in the dark (blue curve) and after illumination at 80 K with 365 nm LED (red curve).

Close modal

Two mechanisms produce such persistent photocapacitance changes at low temperatures. First, if traps with levels below the Fermi level in the space charge region (SCR) are photoionized, they can only return to the steady state by thermal emission of holes, which happens slowly at low temperatures. Second, if deep centers possess a barrier for capture of electrons, these electrons, once removed from the centers by light, can only return to their host centers at temperatures high enough to overcome the barrier for capture. These two situations can be distinguished by applying a forward bias pulse supplying electrons to the space charge region. In the first case, the photogenerated charge can be wiped out by flooding the SCR with electrons during the forward bias pulse. In the second case, the photogenerated charge will not be removed.18,23,24 As seen in Fig. 3(c), with 3.4 eV photon excitation, both situations are encountered: after illumination, the concentration is persistent (the persistent photocapacitance curve, PPC), but can be partially quenched by application of a high forward bias (+2 V), long (2 min) pulse (PPC + 2 V curve). The spectral dependence of concentrations of such PPC and PPC + 2 V centers is shown in Fig. 3(b). For photons with energies >2 eV, centers with a barrier for capture of electrons are excited (all three photogenerated concentrations are close to each other), while, for photons with energies >2.7 eV, there appears a growing number of PPC electrons that can be partially quenched with forward bias application and thus are due to ionization of deep centers in the lower half of the band gap.

Some idea of the temperatures at which the persistent photocapacitance vanishes can be derived from C-T dependences, measured in the dark and after illumination at low temperature. These are compared for sample 2 in Fig. 3(d). The PPC curve obtained after illumination with 365 nm (3.4 eV) LED returns to the dark curve at ∼400 K. A more quantitative estimate of the energy barrier involved can be obtained from ODLTS spectra measurements with optical pulses provided by the 3.4 eV photons. The spectrum for sample 1 is shown in Fig. 4. A wide hole-trap-like ODLTS peak (the capacitance after the excitation pulse decreases with time, a negative peak in the DLTS/ODLTS spectra according to the convention adopted here) is seen. It is comprised of two peaks, with activation energies 0.3 and 0.6 eV. Clearly, 3.4 eV photons cannot excite hole traps near Ev + 0.3 eV or Ev + 0.6 eV in α-Ga2O3 with band gap close to 5 eV. Thus, we assume the transitions in these measurements are due to persistent photoexcitation of deep traps with barriers for capture of electrons ∼0.3 and 0.6 eV (for such centers, the capacitance after excitation will slowly decrease with time, as in the case of true hole traps24).

FIG. 4.

ODLTS spectrum for sample 1 obtained with 365 nm (3.4 eV) LED excitation, reverse bias −1 V, time windows 2.8 s/28 s, pulse length of 5 s.

FIG. 4.

ODLTS spectrum for sample 1 obtained with 365 nm (3.4 eV) LED excitation, reverse bias −1 V, time windows 2.8 s/28 s, pulse length of 5 s.

Close modal

According to spectral measurements of photocapacitance summarized by Fig. 3(c), there should also be deep traps with optical ionization energy >2.7 eV that do not possess a barrier for capture of electrons and can be quenched by the application of forward bias pulse. Such traps could not be detected in our ODLTS spectra measured to 470 K, most likely because of the very high thermal ionization energy of ∼2.3 eV of such centers. This is similar to β-Ga2O3 films grown by HVPE or MOCVD.23,24 In that material, there were also hole traps at Ev + 1.4 eV attributed to Ga vacancies.18,23,24

The spectra of deep electron traps in the upper half of the band gap of the epilayers were studied by DLTS measurements performed at ∼10 kHz, which prevented reliable spectra collection for biases higher than 1 V for sample 2 and higher than 0.5 V for sample 3 because of high leakage current. It also limited the temperatures to which DLTS spectra could be reliably interpreted to 350 K for sample 2 and ∼300 K for sample 3. The most detailed measurements could be done for the lightly doped sample 1. Figure 5(a) shows DLTS spectrum obtained for quiescent bias of −3 V, with voltage pulsed to 0 V for 3 s. The abscissa axis shows the DLTS signal, ΔC/C, multiplied by 2Nd × F−1 where Nd is the shallow donor concentration obtained from C-V, ΔC = C(t1) − C(t2), t1 and t2 are the time windows for processing the DLTS spectra, C is the steady-state capacitance, and F−1 is the spectrometer function.21 For temperatures corresponding to peaks in DLTS spectra, the magnitude of the peak in these coordinates is equal to the concentration of the centers without accounting for the λ-correction.21 

FIG. 5.

(a) DLTS spectra for sample 1 with reverse bias −3 V and forward voltage pulse to 0 V (pulse length 3 s) (red curve), sample 2 at −1 V (blue curve), and sample 3 at −0.5 V (olive line; the amplitudes divided by 10 to bring the data to scale with the two other samples), in all cases time windows were 1.75 s/17.5 s; (b) Arrhenius plots for the A, B, C, D traps.

FIG. 5.

(a) DLTS spectra for sample 1 with reverse bias −3 V and forward voltage pulse to 0 V (pulse length 3 s) (red curve), sample 2 at −1 V (blue curve), and sample 3 at −0.5 V (olive line; the amplitudes divided by 10 to bring the data to scale with the two other samples), in all cases time windows were 1.75 s/17.5 s; (b) Arrhenius plots for the A, B, C, D traps.

Close modal

Two prominent electron traps are present: A (energy Ea = 0.6 eV, electron capture cross section σn = 3 × 10−15 cm2) and B (Ea = 1.1 eV, σn = 1.2 × 10−13 cm2). The concentrations of both traps with account for the λ-correction are given in Table I and are >1014 cm−3. For more heavily doped sample 2, the reverse bias had to be decreased to −1 V, and even so, for temperatures close to 400 K, reliable measurements were not possible. The same A traps, but with a much higher concentration, and additional deep traps C (Ea = 0.85 eV, σn = 1.2 × 10−16 cm2) could be observed. Even for the most heavily doped sample 3, DLTS measurements were possible for temperatures <300 K, although the applied voltage had to be decreased to −0.5 V to avoid excessive leakage. The same A traps, but with a much higher concentration, could be observed. Additionally, another shallow electron traps D (Ea = 0.23 eV, σn = 3 × 10−13 cm2) were detected [the concentrations in Fig. 5(a) are divided by 10 for sample 3 to bring all DLTS spectra to scale]. Figure 5(b) presents the Arrhenius plots of the trap signatures for traps A, B, C, and D detected in our experiments. Table I gives the trap concentrations in sample 3 with account for λ-correction. From Fig. 5 and Table I, it can be seen that increasing the Sn doping steadily increases the concentration of the A electron traps and introduces new traps C and D. The dominant traps A and B and the trap C are quite similar to deep traps observed in semi-insulating α-Ga2O3 grown by pulsed HVPE.14 The energy levels and concentrations of the A and B traps in the lightly doped sample 1 are not very different from the dominant traps in similarly doped β-Ga2O3 grown by HVPE.18,19,24,25,27

The diffusion length of the nonequilibrium charge carriers determined from fitting the EBIC signal collection efficiency Ic/(IbEb) dependence on beam energy Eb16 yields diffusion length Ld = 0.15 µm (Fig. 6), lower than the typical values in HVPE grown β-Ga2O3 (0.4–0.5 µm)16,18,19 (Ic is the EBIC current, and Ib is the probing electron beam current of SEM, 1 nA in this case). Since the dislocation density in our samples is high, it cannot be ruled out that the diffusion length is limited by the dimensions of dislocation bounded cells as is the case for GaN with dislocation densities >5 × 108 cm−2.17 More studies are needed to check whether the diffusion lengths are limited by the density of specific deep traps, as observed for low dislocation density HVPE β-Ga2O3 films.18 DLTS/ODLTS measurements have to be conducted on α-Ga2O3 with various dislocation densities and compared to Ld measurements by EBIC.

FIG. 6.

EBIC collection efficiency as function of beam energy Eb for sample 1 (solid blue squares) and results of fitting with diffusion length Ld = 0.15 µm (magenta line).

FIG. 6.

EBIC collection efficiency as function of beam energy Eb for sample 1 (solid blue squares) and results of fitting with diffusion length Ld = 0.15 µm (magenta line).

Close modal

The evolution of MCL spectra of the α-Ga2O3 with changing Sn doping and temperature were also examined. Earlier publications27,28 describing photoluminescence (PL) or MCL spectra of α-Ga2O3 refer to powder material. The spectra were dominated by a broad band extending from ∼2 eV (625 nm) to ∼3.6 eV (344 nm), with a sharp peak near 3.445 eV (360 nm) whose magnitude increased with decreasing temperature. Figure 7(a) compares the room temperature MCL spectra of samples 1 and 3. These spectra were broad and consisted of a superposition of several peaks. For comparison, we also show in Fig. 7(a) the spectrum taken for high quality undoped semi-insulating layer. Lowering the measurement temperature from 300 to 90 K did not change the peak positions, but increased the MCL intensity and changed the relative intensities of the peaks, suggesting this involves donor-acceptor pairs (DAP) recombination with different shallow and deep donors and deep acceptors [Figs. 7(b) and 7(c) compare the room temperature and 90 K spectra of samples 1 and 3]. Deconvolution of the low temperature MCL spectra of the two samples into several Gaussian peaks yielded the peak positions as 339 nm (3.66 eV), 389 nm (3.19 eV), and 423 nm (2.93 eV) for the low doped sample 1 and 339 nm (3.66 eV), 428 nm (2.89 eV), 441 nm (2.81 eV), 489 nm (2.54 eV), and 540 nm (2.296 eV) for heavily doped sample 3. For the undoped SI sample, the Gaussian bands peaked at 385 nm (3.22 eV), 425 nm (2.92 eV), and 439 nm (2.82 eV). In neither sample did we observe band-edge luminescence. The attribution of the MCL bands to specific transitions cannot be reliably done at this point, but the shortest wavelength band at 339 nm (3.66 eV) is only observed for the two Sn-doped samples and may be related to DAP transitions involving Sn donors. The bands peaked near 385–389 nm (∼3.2 eV) and 423–428 nm (∼2.9 eV) are common for all samples, and the band near 439–441 nm (∼2.8 eV) is observed for heavily Sn-doped sample 3 and for the SI sample, while the long-wavelength band near 540 nm (2.3 eV) is specific to highly Sn-doped sample 3. Similar broad MCL spectra are often observed for β-Ga2O3 films and attributed to recombination involving polaronic hole states in the valence band and different kinds of DAP transitions.29,30

FIG. 7.

(a) Room temperature MCL spectra of sample 1 (1017 cm−3) (blue curve) and sample 3 (4.8 × 1019 cm−3) (magenta curve); for comparison, the spectrum for the undoped high-quality semi-insulating sample is shown as an olive curve; (b) room temperature and 90 K spectra for sample 1; (c) room temperature and 90 K spectra for sample 3.

FIG. 7.

(a) Room temperature MCL spectra of sample 1 (1017 cm−3) (blue curve) and sample 3 (4.8 × 1019 cm−3) (magenta curve); for comparison, the spectrum for the undoped high-quality semi-insulating sample is shown as an olive curve; (b) room temperature and 90 K spectra for sample 1; (c) room temperature and 90 K spectra for sample 3.

Close modal

The concentrations of deep traps in Sn-doped n-type α-Ga2O3 are the same order of magnitude as for β-Ga2O3 films and crystals,18,19,24–26 suggesting electron trapping-detrapping and current collapse in α-Ga2O3 rectifiers and FETs are not expected to be more severe than in β-Ga2O3. The trap levels also scale with those in the β-polytype: the concentrations of the dominant A traps increases for increased donor doping, similar to the E1 electron traps in β-Ga2O3.19,25,26 Heavy doping that promotes point defect formation in α-Ga2O3 generates shallower electron traps D, similar to the effect of proton irradiation for β-Ga2O3 films grown by HVPE.18 The optical ionization thresholds of the deep acceptors detected in photocapacitance in α-Ga2O3 are considerably higher than in β-Ga2O3,18,24 which probably reflects the higher band gap of α-Ga2O3.

The work at NUST MISiS was supported in part by Ministry of Education and Science of the Russian Federation in the framework of Increase Competitiveness Program of NUST MISiS (No. K2-2017-068). The samples were grown by Perfect Crystals LLC. The work at IMT RAS was supported by the State Task No. 075-00475-19-00. The work at UF was sponsored by Department of the Defense, Defense Threat Reduction Agency, No. HDTRA1-17-1-011, and monitored by Jacob Calkins and also by No. NSF DMR 1856662 (Tania Paskova).

1.
S. I.
Stepanov
,
V. I.
Nikolaev
,
V. E.
Bougrov
, and
A. E.
Romanov
,
Rev. Adv. Mater. Sci.
44
,
63
(
2016
), http://www.ipme.ru/e-journals/RAMS/no_14416/06_14416_stepanov.pdf.
2.
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
).
3.
Z.
Galazka
,
Semicond. Sci. Technol.
33
,
113001
(
2018
).
4.
S. J.
Pearton
,
F.
Ren
,
M.
Tadjer
, and
J.
Kim
,
Appl. Phys.
124
,
220901
(
2018
).
5.
M. A.
Mastro
,
A.
Kuramata
,
J.
Calkins
,
J.
Kim
,
F.
Ren
, and
S. J.
Pearton
,
ECS J. Solid State Sci. Technol.
6
,
P356
(
2017
).
6.
N.
Ueda
,
H.
Hosono
,
R.
Waseda
, and
H.
Kawazoe
,
Appl. Phys. Lett.
71
,
933
(
1997
).
7.
M.
Schubert
,
R.
Korlacki
,
S.
Knight
,
T.
Hofmann
,
S.
Schöche
,
V.
Darakchieva
 et al.,
Phys. Rev. B
93
,
125209
(
2016
).
8.
A.
Fiedler
,
R.
Schewski
,
Z.
Galazka
, and
K.
Irmscher
,
ECS J. Solid State Sci. Technol.
8
,
Q3083
(
2019
).
9.
M. J.
Tadjer
,
L. E.
Luna
,
E. R.
Cleveland
,
K. D.
Hobart
, and
F. J.
Kub
,
ECS Trans.
85
,
21
(
2018
).
10.
K.
Kaneko
,
S.
Fujita
, and
T.
Hitora
,
Jpn. J. Appl. Phys., Part 2
57
,
02CB18
(
2018
).
11.
M.
Oda
,
R.
Tokuda
,
H.
Kambara
,
T.
Tanikawa
,
T.
Sasaki
, and
T.
Hitora
,
Appl. Phys. Express
9
,
021101
(
2016
).
12.
Y.
Oshima
,
K.
Kawara
,
T.
Shinohe
,
T.
Hitora
,
M.
Kasu
, and
S.
Fujita
,
APL Mater.
7
,
022503
(
2019
).
13.
A. I.
Pechnikov
,
S. I.
Stepanov
,
A. V.
Chikiryaka
,
M. P.
Shcheglov
,
M. A.
Odnoblyudov
, and
V. I.
Nikolaev
, “
Thick α-Ga2O3 films grown on sapphire by halide vapor phase epitaxy
,”
Russ. Phys. Semicond.
6
,
789
(
2019
) (in Russian).
14.
D.-W.
Jeon
,
H.
Son
,
J.
Hwang
,
A. Y.
Polyakov
,
N. B.
Smirnov
,
I. V.
Shchemerov
,
A. V.
Chernykh
,
A. I.
Kochkova
,
S. J.
Pearton
, and
In-H.
Lee
,
APL Mater.
6
,
121110
(
2018
).
15.
A. Y.
Polyakov
,
H.-S.
Cho
,
J.-H.
Yun
,
In-H.
Lee
,
E. B.
Yakimov
,
N. B.
Smirnov
, and
K. D.
Shcherbachev
,
ECS J. Solid State Sci. Technol.
5
,
Q165
(
2016
).
16.
E. B.
Yakimov
,
A. Y.
Polyakov
,
N. B.
Smirnov
,
I. V.
Shchemerov
,
J.
Yang
,
F.
Ren
,
G.
Yang
,
J.
Kim
, and
S. J.
Pearton
,
J. Appl. Phys.
123
,
185704
(
2018
).
17.
E. B.
Yakimov
,
A. Y.
Polyakov
,
In-H.
Lee
, and
S. J.
Pearton
, “
Recombination properties of dislocations in GaN
,”
J. Appl. Phys.
123
,
161543
(
2018
).
18.
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
).
19.
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
).
20.
L.
Pautrat
,
B.
Katircioglu
,
N.
Magnea
,
D.
Bensahel
,
J. C.
Pfister
, and
L.
Revoil
,
Solid-State Electron.
23
,
1159
(
1980
).
21.
D. K.
Schroder
, in
Semiconductor Materials and Device Characterization
(
John Wiley & Sons
,
New York
,
1990
), Chap. 7.
22.
A. Y.
Polyakov
,
N. B.
Smirnov
,
In-H.
Lee
, and
S. J.
Pearton
,
J. Vac. Sci. Technol., B: Nanotechnol. Microelectron.: Mater., Process., Meas., Phenom.
33
,
061203
(
2015
).
23.
A. Y.
Polyakov
,
N. B.
Smirnov
,
I. V.
Shchemerov
,
D.
Gogova
,
S. A.
Tarelkin
, and
S. J.
Pearton
,
J. Appl. Phys.
123
,
115702
(
2018
).
24.
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
).
25.
M. E.
Ingebrigtsen
,
J. B.
Varley
,
A. Yu.
Kuznetsov
,
B. G.
Svensson
,
G.
Alfieri
,
A.
Mihaila
,
U.
Badstübner
, and
L.
Vines
,
Appl. Phys. Lett.
112
,
042104
(
2018
).
26.
E.
Farzana
,
E.
Ahmadi
,
J. S.
Speck
,
A. R.
Arehart
, and
S. A.
Ringel
,
J. Appl. Phys.
123
,
161410
(
2018
).
27.
S.
Cho
,
J.
Lee
,
In Y.
Park
, and
S.
Kim
,
Jpn. J. Appl. Phys., Part 1
41
,
5237
(
2002
).
28.
H. W.
Leverenz
,
An Introduction to Luminescence of Solids
(
Dover
,
New York
,
1968
), p.
222
.
29.
T.
Onuma
,
Y.
Nakata
,
K.
Sasaki
,
T.
Masui
,
T.
Yamaguchi
,
T.
Honda
,
A.
Kuramata
,
S.
Yamakoshi
, and
M.
Higashiwaki
,
J. Appl. Phys.
124
,
075103
(
2018
).
30.
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
).