The impact of electron injection, using 10 keV beam of a Scanning Electron Microscope, on minority carrier transport in Si-doped β-Ga2O3 was studied for temperatures ranging from room to 120°C. In-situ Electron Beam-Induced Current technique was employed to determine the diffusion length of minority holes as a function of temperature and duration of electron injection. The experiments revealed a pronounced elongation of hole diffusion length with increasing duration of injection. The activation energy, associated with the electron injection-induced elongation of the diffusion length, was determined at ∼ 74 meV and matches the previous independent studies. It was additionally discovered that an increase of the diffusion length in the regions affected by electron injection is accompanied by a simultaneous decrease of cathodoluminescence intensity. Both effects were attributed to increasing non-equilibrium hole lifetime in the valence band of β-Ga2O3 semiconductor.

Wide bandgap semiconductors such as GaN and ZnO are important for a number of applications ranging from flame detection and high temperature electronics to solar-blind ultraviolet detectors and sensors for use in extreme environmental conditions.1–3 An emerging semiconductor with a direct bandgap of 4.69 eV (ultra-wide bandgap), Ga2O3 has become an attractive candidate for radiation applications in terms of its superior stability over GaN and ZnO.1,4–9

So far, the main limiting factor in Ga2O3 technology is related to difficulties of p-type conductivity realization.7,9–11 Previous studies indicate that holes in β-Ga2O3 have characteristics of low dispersion, high effective mass and high density of states. This results in formation of weak polarons (or localized holes), trapped by lattice distortions in the vicinity.12–16 Contradicting these claims, Refs. 6 and 10 has recently experimentally shown a possibility of p-type conductivity in Ga2O3. It is additionally reported that the self-trapping nature of holes is destroyed at temperatures above 90-120 K.11 Given the current state-of-the-art and ever mounting interest in p-type Ga2O3, one can reasonably claim that a robust p-type Ga2O3 electrical conductivity, if not yet fully feasible, will be possible in the foreseeable future.

In Ga2O3, if bipolar devices become available, minority carrier transport (diffusion length) will be of primary importance. Minority carrier diffusion length defines the performance of bipolar devices such as p-n junction diodes, bipolar transistors and p-i-n detectors.6,9,17 One of the major issues in the current ZnO and GaN device technology, is the low diffusion length of minority carriers, partially due to dislocation scattering.18,19 It has been previous shown that in p-type GaN and ZnO, electron injection either with an electron beam or forward bias, results in significant increase in minority carrier diffusion length.18,20–23 Such an increase in the diffusion length translates into changes in the material’s optical properties and improved photovoltaic detector performance.

Recent studies on diffusion length of minority holes in n-type β-Ga2O3 report the value ∼ 400 nm at room temperature.24,25 Hence, like in the case of ZnO and GaN, the issue of short diffusion length for minority carriers also exists in β-Ga2O3 and limits its possible applications in bipolar devices. This paper demonstrates the impact of electron injection on minority carrier transport in Si-doped β-Ga2O3. This study testifies that: 1) one can engineer minority carrier diffusion length with electron injection (increase it several times); 2) the non-equilibrium holes, generated due to electron beam, are not self-trapped and contribute profoundly to the Electron Beam-Induced Current.

Electron Beam-Induced Current (EBIC) and Cathodoluminescence (CL) techniques3,18,20–23 were employed in-situ in Philips XL30 Scanning Electron Microscope (SEM) to study the effect of electron injection on minority carrier transport, namely, diffusion length (L) and lifetime (τ) in β-Ga2O3. The samples involved in the study consisted of epitaxial Si-doped β-Ga2O3 grown on Sn-doped β-Ga2O3 substrate. The epitaxial layer growth was carried out by Halide Vapor-Phase Epitaxy (HVPE), courtesy Novel Crystal Technology, and the Sn-doped β-Ga2O3 substrate with [001] orientation was grown by edge-defined film-fed method (Tamura Corporation, Japan). Hall measurements on the n+- doped substrate showed a carrier concentration of ∼ 2.2×1018 cm-3. The initial thickness of epitaxially grown Si-doped (with electron concentration ∼ 3.6×1016 cm-324) β-Ga2O3 layer was 20 μm, which was subsequently reduced to 10 μm after chemical-mechanical polishing. Schottky rectifiers were obtained by making Schottky contacts (Ni – 20 nm/Au – 80 nm) on the epitaxial Si-doped β-Ga2O3 layer with photolithography/liftoff technique, and Ohmic contact on the Sn-doped β-Ga2O3 substrate (Ti – 20 nm/Au – 80nm) by blanket evaporation over the entire back side of the substrate. The diameter of the top contacts, used for measurements, was 210 μm.

The diffusion length of minority carriers was determined using line-scan EBIC technique in the planar configuration24,26–30 with electron beam energy of 10 keV (corresponding to the electron range of ∼ 0.4 μm and the absorbed current of ∼ 0.4 nA in the material) in the temperature range of 23-120 °C (hot-stage of Gatan MonoCL2 system, integrated with Philips SEM, was used for varying temperature). A single line-scan, carried out from the edge of the Schottky barrier outwards (cf. Fig. 1, top-right inset), takes about 6 seconds. While L values can be extracted from this single line-scan, in this work, a scanning of the region of interest was not interrupted and was continued for a total duration of electron injection up to 720 seconds with periodic L measurements. Since this material was very sensitive to electron injection, the measurements for various temperatures were done at different locations, to exclude any possibility of uncalibrated electron injection. Line-scans were recorded using a home-made software interfaced with Stanford Research Systems Low-Noise Current Amplifier (SR570) and Keithley DMM 2000 digital multimeter.

FIG. 1.

A sample EBIC line-scan (I vs. distance) taken at room temperature. The distance is measured from the edge of the Schottky barrier and the fit is determined according to the equation (1). Inset (top-right): Schematic diagram of the measurement setup, (bottom-left): Linear fit according to the equation (2).

FIG. 1.

A sample EBIC line-scan (I vs. distance) taken at room temperature. The distance is measured from the edge of the Schottky barrier and the fit is determined according to the equation (1). Inset (top-right): Schematic diagram of the measurement setup, (bottom-left): Linear fit according to the equation (2).

Close modal

Variable-temperature CL measurements were carried out using Gatan MonoCL2 attachment to the SEM and an integrated temperature controller with an accuracy of 0.5 °C. The CL spectra were recorded with a single grating (1200 lines/mm blazed at 500 nm) monochromator and a Hamamatsu photomultiplier tube, sensitive in the 150-850 nm range. SEM electron beam with energy of 10 keV was used for excitation of 1 μm x 1 μm area, located at 10 μm distance from the Schottky contact, for a total electron injection duration of 3000 seconds.

The material was first analyzed at “zero” electron injection dose before studying the impact of electron injection. A single line-scan of 6 seconds was used, which had virtually no impact on L. A typical EBIC line-scan taken at room temperature is shown in Fig. 1, where the current is shown as a function of beam position measured from the edge of the Schottky contact.

L was extracted from the experimental data in Fig. 1 using the following equation:27,29,31–33

(1)

where, x is the beam-to-junction distance, I0 is a scaling constant, L is the diffusion length of minority carriers (holes, in this case), and α represents a surface recombination velocity and takes values between -1/2 and -3/2. Equation (1) is true for x>>d, where d is the depletion region width. Width of the depletion region is calculated as d = [2εεr(Vbi-V)/(qNB) ]0.5, where εr is the relative permittivity, V is the applied voltage, Vbi is the Schottky barrier built-in potential, q is the electron charge, and NB is the bulk doping level in β-Ga2O3. In this case, εr=10.5,1 Vbi=1.025 V,34 V=0 V, and NB=3.6×1016 cm-3, resulting in d ≈ 179 nm, which is at least 8 times less than x.

In equation (1), I0, α, and L are fit parameters, and the fit was performed at the knee region of the curve in Fig. 1, which ensures the accuracy of the extracted parameters according to the constraint x > 2L.31,35 Moreover, a modified form of equation (1), given below, reduces the three-parameter nonlinear fit to a linear one, as shown in the bottom-left inset of Fig. 1:

(2)

L was calculated as the inverse slope of the linear fit according to equation (2). The best linear fit was obtained for α = -3/2, which is, therefore, used throughout the article.

Temperature dependence of the L is given by:32 

(3)

where k is the Boltzmann constant, L0 is the asymptotic diffusion length, and T is the temperature in Kelvin. ΔEA,T is the activation energy associated with a charge trap in the forbidden gap.32 Temperature dependence of L is shown in Fig. 2. Similar to the previously reported results,25 L exhibits a decrease from ∼ 430 nm at room temperature to ∼ 320 nm at 120 °C.

FIG. 2.

Dependence of hole diffusion length on temperature for a single EBIC line-scan. Inset: The slope of the linear fit using equation (3) gives the thermal activation energy ΔEA,T for the temperature dependence of diffusion length.

FIG. 2.

Dependence of hole diffusion length on temperature for a single EBIC line-scan. Inset: The slope of the linear fit using equation (3) gives the thermal activation energy ΔEA,T for the temperature dependence of diffusion length.

Close modal

The decrease in L with temperature is not commonly observed in other semiconductors. In GaAs, ZnO and GaN, for example, L increases with temperature.3,18,31 The activation energy, estimated from the Arrhenius plot (Fig. 2, inset) using equation (2), is ∼ 53 meV. This energy is most likely associated with Si-donors, which form fairly shallow 50 meV levels.36,37 A likely cause for L reduction with growing temperature could be related to thermalization of carriers.38 Carrier-carrier interaction significantly contributes to thermalization only for high carrier concentrations (> 1x1018 cm-3). For elevated temperatures and relatively low doping levels, such as in this case, carrier-phonon interaction becomes a dominating factor,39,40 which leads to L decrease.

Continuous exposure of the sample to electron beam resulted in a significant x4.5 increase in L as shown in Fig. 3. A similar effect was also observed in p-type ZnO and GaN as reported previously.18,21Fig. 3 (top inset) shows the EBIC line-scans for repeated exposures up to 720 seconds. A clear elongation (tail) is observed in the EBIC signal decay (compare a line-scan after 120 and 720 seconds of continuous electron beam excitation). As can be seen from equation (1), a longer tail in the EBIC signal dependence on distance, manifests in longer minority carrier diffusion length. A marked increase in the amplitude of the EBIC signal is also observed at the same time (cf. top and bottom insets of Fig. 3). The increase in the EBIC signal amplitude is attributed to the rise of the collection efficiency of minority carriers: longer L results in larger number of non-equilibrium minority electrons swept by built-in field of the Schottky barrier used in the measurements. The EBIC collection efficiency increases up to a certain value of L, after which any increase in L (which slows down with electron injection duration; cf. Fig.3, top inset) has no effect on the EBIC signal amplitude.41–43 The diffusion length in Fig. 3 exhibits a saturation (not shown on the plot), following a saturation of Imax (cf. Fig. 3, bottom inset). It should be noted that the increased values of L, achieved due to electron beam injection, persist for at least one day at room temperature following injection.

FIG. 3.

Linear increase in the diffusion length due to continuous electron beam irradiation. Inset (top): Electron Beam Induced-Current vs. distance line-scans at various durations of electron injection. Inset (bottom): EBIC signal amplitude as a function of electron injection duration.

FIG. 3.

Linear increase in the diffusion length due to continuous electron beam irradiation. Inset (top): Electron Beam Induced-Current vs. distance line-scans at various durations of electron injection. Inset (bottom): EBIC signal amplitude as a function of electron injection duration.

Close modal

Fig. 4 shows the effect of moderate electron injection duration on minority carrier diffusion length. This study was carried out at variable temperature to understand its impact on the rate of L increase (μm/sec) due to electron injection. The plot for room temperature in Fig. 4 can be correlated with the 0-100 second segment of L versus electron injection duration in Fig. 3, with a slight mismatch in slope attributed to different locations for two experiments.

FIG. 4.

Temperature dependence for minority hole diffusion length increase due to electron injection. The rate, R, of diffusion length elongation decreases with temperature. Inset: The slope of the linear fit with equation (4) giving the activation energy ΔEA,I for the electron injection-induced increase of L.

FIG. 4.

Temperature dependence for minority hole diffusion length increase due to electron injection. The rate, R, of diffusion length elongation decreases with temperature. Inset: The slope of the linear fit with equation (4) giving the activation energy ΔEA,I for the electron injection-induced increase of L.

Close modal

The effect of temperature on the rate, R, of L increase with electron injection duration can be described as:18 

(4)

where R0 is a constant, ΔEA,T is thermal activation energy for diffusion length (from equation (3)), and ΔEA,I is the activation energy due to electron injection effect.

As R implicitly contains the injection information in the form of L versus electron injection duration, t (R=dL/dt), use of equation (4) to calculate ΔEA,I is justified. In equation (4), the first exponent represents L dependence on temperature (cf. equation (3), in which a ½ factor in the exponent is due to equation (5) (see below)); the second exponent is a Boltzmann factor, which is used to model the temperature dependence of electron injection effect. One can see from Fig. 4 that R decreases with increasing temperature. Fig. 4 (inset) shows the Arrhenius plot with a linear fit of R vs. 1/kT, according to equation (4), yielding ΔEA,I to be ∼ 74 meV. Similar activation energy, attributed to yet unknown recombination center, was recently observed in an independent study.44 

The diffusion length of minority carriers is an integral quantity, which is related to the lifetime of non-equilibrium carriers in the band by the relation:19 

(5)

where D is the diffusion coefficient and τ is the minority carrier lifetime. A larger lifetime of non-equilibrium carriers in the band indicates lower number of radiative recombination events.45 If L increase is due to an increase in τ, the CL intensity from the radiative recombination process should decrease with duration of electron injection. Indeed, from the CL experiment performed at room temperature (Fig. 5), a decrease in the CL intensity I, was observed with the duration of electron injection.

FIG. 5.

CL spectra from the region affected by electron injection as a function of its duration. Inset: Dependence of the square root for normalized (relative to the maximum intensity I0) inversed intensity on duration of electron injection, and the linear fit.

FIG. 5.

CL spectra from the region affected by electron injection as a function of its duration. Inset: Dependence of the square root for normalized (relative to the maximum intensity I0) inversed intensity on duration of electron injection, and the linear fit.

Close modal

Earlier studies have shown that I1τ.45 Moreover, L depends linearly on electron injection duration (Figs. 3, 4) and on τ (equation (5)). Hence, it is concluded that the dependence of CL intensity on electron injection duration, t, should be given by I1/2t.45 This relation was verified by the linear fit in Fig. 5, inset, which is also in agreement with the assumption that the linear increase of L with injection duration is due to an increased lifetime for minority carriers.

The proposed model for the observed electron injection-induced effects in β-Ga2O3 is as follows:

  • Non-equilibrium electron-hole pairs are generated in β-Ga2O3 due to electron beam irradiation. The total charge, injected into the region under investigation due to this irradiation, is up to 1.2 μC.

  • We expect that the unknown traps with activation energy of 74 meV (see above) can capture some non-equilibrium electrons. A similar situation was observed in Mg-doped GaN and Sb-doped ZnO.3,18,21–23

  • As a portion of non-equilibrium electrons is captured, these traps stop playing a role in the recombination process (mostly radiative, cf. Fig. 5) and the non-equilibrium electrons and holes live longer life (τ) in the conduction and valence bands, respectively. Increased τ leads to longer minority carrier diffusion length L (as measured by EBIC) and suppressed CL intensity from the region subjected to electron irradiation (cf. Fig. 5). As the electron injection continues, the number of traps, available to capture non-equilibrium electrons, saturates. As a result, the EBIC signal amplitude (cf. insets of Fig. 3) and the CL intensity (cf. Fig. 5) reach their saturation levels.

  • As the temperature increases, trapped electrons have a higher chance to escape and recombine with holes in the valence band. At the same time, the traps again become available for recombination. This leads to a reduced rate R (for L increase with electron irradiation duration) at elevated temperatures (cf. Fig. 4).

In conclusion, this study demonstrates a significant increase of minority hole diffusion length in n-type β-Ga2O3. Enhancement of minority carrier transport is accompanied by the consistent changes in the material’s optical properties. Both effects are related to an increased lifetime for non-equilibrium carriers in the conduction and valence bands (non-equilibrium electrons and holes are generated in pairs). Additionally, the experiments show that minority holes may substantially contribute to current in β-Ga2O3 semiconductor, which is an argument against self-trapping at room and elevated temperatures.

Research at UCF and the Weizmann institute was supported in part by NATO (award # G5453) and NSF (UCF award # ECCS1802208). The work at UF was sponsored by the Department of Defense, Defense Threat Reduction Agency, HDTRA1-17-1-011, monitored by Jacob Calkins.

1.
S. J.
Pearton
,
J.
Yang
,
P. H.
Cary
,
F.
Ren
,
J.
Kim
,
M. J.
Tadjer
, and
M. A.
Mastro
,
Applied Physics Reviews
5
,
011301
(
2018
).
2.
J. V.
Li
,
S. L.
Chuang
,
E. M.
Jackson
, and
E.
Aifer
,
Appl. Phys. Lett.
85
,
1984
(
2004
).
3.
L.
Chernyak
,
A.
Osinsky
, and
A.
Schulte
,
Solid-State Electron.
45
,
1687
(
2001
).
4.
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
).
5.
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
).
6.
J.
Yang
,
F.
Ren
,
S. J.
Pearton
,
G.
Yang
,
J.
Kim
, and
A.
Kuramata
,
Journal of Vacuum Science & Technology B, Nanotechnology and Microelectronics: Materials, Processing, Measurement, and Phenomena
35
,
031208
(
2017
).
7.
A.
Mock
,
R.
Korlacki
,
C.
Briley
,
V.
Darakchieva
,
B.
Monemar
,
Y.
Kumagai
,
K.
Goto
,
M.
Higashiwaki
, and
M.
Schubert
,
Phys. Rev. B
96
,
245205
(
2017
).
8.
M. A.
Mastro
,
A.
Kuramata
,
J.
Calkins
,
J.
Kim
,
F.
Ren
, and
S. J.
Pearton
,
ECS Journal of Solid State Science and Technology
6
,
P356
(
2017
).
9.
S. I.
Stepanov
,
V. I.
Nikolaev
,
V. E.
Bougrov
, and
A. E.
Romanov
,
Rev. Adv. Mater. Sci.
44
,
63
(
2016
).
10.
E.
Chikoidze
,
A.
Fellous
,
A.
Perez-Tomas
,
G.
Sauthier
,
T.
Tchelidze
,
C.
Ton-That
,
T. T.
Huynh
,
M.
Phillips
,
S.
Russell
,
M.
Jennings
,
B.
Berini
,
F.
Jomard
, and
Y.
Dumont
,
Materials Today Physics
3
,
118
(
2017
).
11.
B. E.
Kananen
,
N. C.
Giles
,
L. E.
Halliburton
,
G. K.
Foundos
,
K. B.
Chang
, and
K. T.
Stevens
,
J. Appl. Phys.
122
,
215703
(
2017
).
12.
J. B.
Varley
,
A.
Janotti
,
C.
Franchini
, and
C. G.
Van de Walle
,
Phys. Rev. B
85
(
2012
).
13.
S.
Yoshioka
,
H.
Hayashi
,
A.
Kuwabara
,
F.
Oba
,
K.
Matsunaga
, and
I.
Tanaka
,
J. Phys.: Condens. Matter
19
,
346211
(
2007
).
14.
H.
He
,
R.
Orlando
,
M. A.
Blanco
,
R.
Pandey
,
E.
Amzallag
,
I.
Baraille
, and
M.
Rérat
,
Phys. Rev. B
74
,
195123
(
2006
).
15.
H.
He
,
M. A.
Blanco
, and
R.
Pandey
,
Appl. Phys. Lett.
88
,
261904
(
2006
).
16.
J.
Furthmüller
and
F.
Bechstedt
,
Phys. Rev. B
93
,
115204
(
2016
).
17.
M.
Higashiwaki
,
K.
Konishi
,
K.
Sasaki
,
K.
Goto
,
K.
Nomura
,
Q. T.
Thieu
,
R.
Togashi
,
H.
Murakami
,
Y.
Kumagai
,
B.
Monemar
,
A.
Koukitu
,
A.
Kuramata
, and
S.
Yamakoshi
,
Appl. Phys. Lett.
108
,
133503
(
2016
).
18.
O.
Lopatiuk-Tirpak
,
L.
Chernyak
,
F. X.
Xiu
,
J. L.
Liu
,
S.
Jang
,
F.
Ren
,
S. J.
Pearton
,
K.
Gartsman
,
Y.
Feldman
,
A.
Osinsky
, and
P.
Chow
,
J. Appl. Phys.
100
,
086101
(
2006
).
19.
L.
Chernyak
,
A.
Osinsky
,
G.
Nootz
,
A.
Schulte
,
J.
Jasinski
,
M.
Benamara
,
Z.
Liliental-Weber
,
D. C.
Look
, and
R. J.
Molnar
,
Appl. Phys. Lett.
77
,
2695
(
2000
).
20.
L.
Chernyak
,
G.
Nootz
, and
A.
Osinsky
,
Electron. Lett.
37
,
922
(
2001
).
21.
L.
Chernyak
,
A.
Osinsky
,
V.
Fuflyigin
, and
E. F.
Schubert
,
Appl. Phys. Lett.
77
,
875
(
2000
).
22.
O.
Lopatiuk-Tirpak
,
L.
Chernyak
,
Y. L.
Wang
,
F.
Ren
,
S. J.
Pearton
,
K.
Gartsman
, and
Y.
Feldman
,
Appl. Phys. Lett.
90
,
172111
(
2007
).
23.
O.
Lopatiuk-Tirpak
,
L.
Chernyak
,
L. J.
Mandalapu
,
Z.
Yang
,
J. L.
Liu
,
K.
Gartsman
,
Y.
Feldman
, and
Z.
Dashevsky
,
Appl. Phys. Lett.
89
,
142114
(
2006
).
24.
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
).
25.
J.
Lee
,
E.
Flitsiyan
,
L.
Chernyak
,
J.
Yang
,
F.
Ren
,
S. J.
Pearton
,
B.
Meyler
, and
Y. J.
Salzman
,
Appl. Phys. Lett.
112
,
082104
(
2018
).
26.
O.
Kurniawan
and
V. K. S.
Ong
,
2006 Conference on Optoelectronic and Microelectronic Materials and Devices
,
IEEE
,
Perth, WA, Australia
(
2006
).
27.
D. E.
Ioannou
and
C. A.
Dimitriadis
,
IEEE Trans. Electron Devices
29
,
445
(
1982
).
28.
C. A.
Dimitriadis
,
Journal of Physics D-Applied Physics
14
,
2269
(
1981
).
29.
F.
Berz
and
H. K.
Kuiken
,
Solid-State Electron.
19
,
437
(
1976
).
30.
L.
Jastrzebski
,
J.
Lagowski
, and
H. C.
Gatos
,
Appl. Phys. Lett.
27
,
537
(
1975
).
31.
L.
Chernyak
,
A.
Osinsky
,
H.
Temkin
,
J. W.
Yang
,
Q.
Chen
, and
M.
Asif Khan
,
Appl. Phys. Lett.
69
,
2531
(
1996
).
32.
M.
Eckstein
and
H. U.
Habermeier
,
J. Phys. IV
1
,
23
(
1991
).
33.
H. J.
Leamy
,
J. Appl. Phys.
53
,
R51
(
1982
).
34.
J.
Yang
,
C.
Fares
,
R.
Elhassani
,
F.
Ren
,
S. J.
Pearton
,
M.
Tadjer
, and
A.
Kuramata
, Unpublished Results (
2018
).
35.
D. S. H.
Chan
,
V. K. S.
Ong
, and
J. C. H.
Phang
,
IEEE Trans. Electron Devices
42
,
963
(
1995
).
36.
H.
Masataka
,
K.
Akito
,
M.
Hisashi
, and
K.
Yoshinao
,
J. Phys. D: Appl. Phys.
50
(
2017
).
37.
A. T.
Neal
,
S.
Mou
,
S.
Rafique
,
H.
Zhao
,
E.
Ahmadi
,
J. S.
Speck
,
K. T.
Stevens
,
J. D.
Blevins
,
D. B.
Thomson
,
N.
Moser
,
K. D.
Chabak
, and
G. H.
Jessen
,
Appl. Phys. Lett.
113
,
062101
(
2018
).
38.
P.
Yu
and
M.
Cardona
,
Fundamentals of Semiconductors: Physics and Materials Properties
(
Springer Berlin Heidelberg
,
2010
).
39.
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
).
40.
N.
Ma
,
N.
Tanen
,
A.
Verma
,
Z.
Guo
,
T.
Luo
,
H.
Xing
, and
D.
Jena
,
Appl. Phys. Lett.
109
,
212101
(
2016
).
41.
M.
Ledra
and
N.
Tabet
,
Superlattices and Microstructures
45
,
444
(
2009
).
42.
E. B.
Yakimov
,
S. S.
Borisov
, and
S. I.
Zaitsev
,
Semiconductors
41
,
411
(
2007
).
43.
C.
Donolato
,
Journal de Physique Colloques
50
,
57
(
1989
).
44.
T. T.
Huynh
,
L. L. C.
Lem
,
A.
Kuramata
,
M. R.
Phillips
, and
C.
Ton-That
,
Physical Review Materials
2
(
2018
).
45.
L.
Chernyak
,
W.
Burdett
,
M.
Klimov
, and
A.
Osinsky
,
Appl. Phys. Lett.
82
,
3680
(
2003
).