The use of ITO interlayers between Ga2O3 and Ti/Au metallization is shown to produce Ohmic contacts after annealing in the range of 500–600 °C. Without the ITO, similar anneals do not lead to linear current–voltage characteristics. Transmission line measurements were used to extract the specific contact resistance of the Au/Ti/ITO/Ga2O3 stacks as a function of annealing temperature. Sheet, specific contact, and transfer resistances all decreased sharply from as-deposited values with annealing. The minimum transfer resistance and specific contact resistance of 0.60 Ω mm and 6.3 × 10−5 Ω cm2 were achieved after 600 °C annealing, respectively. The conduction band offset between ITO and Ga2O3 is 0.32 eV and is consistent with the improved electron transport across the heterointerface.

There is renewed interest in Ga2O3 because of its large bandgap, availability of large diameter, high quality crystals, and excellent electrical transport properties. The β-polytype of Ga2O3 in particular can be used to fabricate gas sensors, high power electronic devices such as rectifiers and transistors, and truly solar blind UV photodetectors.1–22 Ga2O3 has a very high theoretical breakdown field (∼8 MV cm−1),1,3,4 making it a candidate for ultrahigh power electronics.18–21 While its thermal conductivity is modest, thermal management techniques developed for GaN will mitigate some of these issues.23–26 

Low resistance Ohmic contacts are a prerequisite for any device and should provide low contact resistance at moderate anneal temperatures. Additional contact resistance leads to slower device switching speeds as well as reliability issues due to local contact heating during current flow during device operation.27–30 The usual approaches to improving contact resistance are to locally increase the doping under the contact through ion implantation or deliberate introduction of vacancy-related donors, or to add a low resistance interlayer. The lowest specific contact resistances reported on Ga2O3 are ∼5 × 10−6 Ω cm−2 for Ti/Au contacts on Si-implanted n-Ga2O3 epitaxial layers.27 Sputtered indium tin oxide (ITO) on n-Ga2O3, followed by annealing at 900–1150 °C (Ref. 28) was also found to improve Ohmic contacts. Pt/ITO contacts on n-Ga2O3 showed superior Ohmic contacts to Pt/Ti and this was attributed to the formation of an interfacial layer with lower bandgap and higher doping concentration than the Ga2O3 alone.29 The band alignment at the heterointerface is also critically important in determining the favorability of carrier transport.30,31 Several authors have found that the presence of upward band bending in low conductivity Ga2O3 complicates achievement of Ohmic contacts.32–37 The surface states present on the Ga2O3 are crucial to the type of contact formed with Au/Ti electrodes. If there are a high density of surface states present, carriers can tunnel through the barrier easily. As a result, a good Ohmic contact between Au/Ti and Ga2O3 will be formed.38 If there are a low density of surface states on the Ga2O3 surface, it should be a Schottky contact.39 

ITO has excellent conductivity and can be deposited in thin film form by the usual semiconductor physical or vapor deposition methods.40–47 It is widely used as transparent electrodes in both GaN-based light-emitting diodes40 and liquid crystal displays.41–47 It is therefore an attractive candidate for use in contact schemes to Ga2O3.

In this paper, we report on the efficacy of ITO interlayers in promoting Ohmic behavior in Ti/Au contacts on n-type Ga2O3 and measure the contact resistance as a function of postdeposition annealing temperature The minimum specific contact resistance of 6.3 × 10−5 Ω cm2, was achieved after 600 °C annealing. The improvement in electron transport across the contact interface is consistent with the band alignment in the ITO/Ga2O3 heterostructure.

The starting wafers were bulk 2 in. diameter β-Ga2O3 single crystals (∼679 μm thick) with (−201) surface orientation (Tamura Corporation, Japan) grown by the edge defined film-fed growth method. Hall effect measurements show unintentional n-type doping of the samples with an electron concentration of ∼3 ×1017 cm−3. On this doped wafer, Si+ implantation was performed at an energy of 30 keV with a dose of 1×1015 cm−2. The samples were then annealed at 950 °C to activate the implanted ions and increase conductivity in the implanted region.

Cl2/Ar based plasmas were employed to define mesa of linear transmission line method (TLM) testers in a Plasma-Therm Versaline inductively coupled plasma (ICP) system. The 300 W (2 MHz) ICP power and 150 rf (13.56 MHz) chuck power resulted a dc self-bias of −150 V on the sample electrode. Ten nanometers of ITO was deposited by RF magnetron sputtering on the Si implanted β-Ga2O3 at room temperature using a 3-in. diameter composite ITO (In2O3/SnO2 90/10) target. The RF power was 125 W and the working pressure was 5×106 Torr in a pure Ar ambient. The DC bias on the electrode under these conditions is in the range of 30–40 V. The contact metal pads for TLM testers were formed by standard lift-off of E-beam deposited Ti/Au (20 nm/80 nm) metallization. The sheet resistance of the ITO/Ti/Au was 420 mΩ/sq. as measured by four point probe. ITO was wet etched from between contact pads and using 1:1 HCl:DI water. The contact pads are square, with dimension of 100 × 100 μm and gaps are 5, 10, 15, 20, 25, and 30 μm. A SSI Solaris 150 rapid thermal annealing system was used to anneal the contact in the nitrogen ambient for 30 s at either 500 or 600 °C. The ramp-up rate was 45 °C s−1, and the ramp-down rate was 32 °C s−1. Figure 1 shows a schematic of the contact stack structure.

Fig. 1.

(Color online) Schematic of (a) Au/Ti and (b) Au/Ti/ITO contact stack on Si-implanted Ga2O3.

Fig. 1.

(Color online) Schematic of (a) Au/Ti and (b) Au/Ti/ITO contact stack on Si-implanted Ga2O3.

Close modal

A Wentworth automated temperature control chuck was used to perform temperature dependent dc measurements, and chuck temperature was varied from room temperature to 150 °C. Current–voltage (I-V) characteristics were measured with an Agilent 4145B parameter analyzer using Be/Cu probe tips.

Figure 2(a) shows the I-V characteristics of Au/Ti/Ga2O3 contact stacks as a function of annealing temperature. Annealing at temperatures up to 500 °C did not lead to Ohmic behavior. This annealing temperature is not sufficient to displace Si atoms that were incorporated into the crystal structure after Si implantation and thermal activation at 950 °C annealing. By sharp contrast, Fig. 2(b) shows that the addition of the ITO interlayer significantly improved current flow into and through the substrate. Low temperature annealing further improved the Ohmic performance, with compliance of 100 mA reached first by the sample annealed at 600 °C.

Fig. 2.

(Color online) I-V of (a) Au/Ti/Ga2O3 and (b) Au/Ti/ITO/Ga2O3 contact stacks as a function of annealing temperature.

Fig. 2.

(Color online) I-V of (a) Au/Ti/Ga2O3 and (b) Au/Ti/ITO/Ga2O3 contact stacks as a function of annealing temperature.

Close modal

The sheet resistance, specific contact resistance, and transfer resistance were calculated from TLM data.48 A relatively thick titanium layer of 10-nm is used to serve as an adhesion layer for the Au layer on the top of the oxide layer. The Ti layer thickness was not optimized, since the emphasis of this paper is to demonstrate the contact with ITO interfacial layer reducing the annealing temperature and achieve reasonable contact resistivity. A sample output resistance versus gap size from ITO contact stack annealed at 500 °C is shown in Fig. 3, a strong linear dependence (r2 = 0.997) is observed.

Fig. 3.

(Color online) TLM data of Au/Ti/ITO/Ga2O3 contact stack resistance as a function of gap spacing.

Fig. 3.

(Color online) TLM data of Au/Ti/ITO/Ga2O3 contact stack resistance as a function of gap spacing.

Close modal

From data of this type, we were able to extract the sheet resistance, specific contact resistance, and transfer resistance as a function of annealing temperature. All of these measurements were performed at room temperature. Specific contact and transfer resistances all decreased sharply from as-deposited values with annealing, as shown in Fig. 4. The lowest temperature data point is unannealed, but was shown to demonstrate the need for annealing to improve contact performance. Samples at 300 °C and 400 °C without mesa definition had been tested prior to this study, but true Ohmic performance was not achieved at those temperatures. Sheet resistance slightly increased with annealing from the as-deposited state due to removal of ion bombardment damage from ITO sputtering. The lowest transfer resistance and specific contact resistance of 0.6 Ω mm and 6.3 × 10−5 Ω cm2 were achieved after 600 °C annealing, respectively. Bae et al.49 reported that at the interface between Ga2O3 and Ti/Au contact layers, annealing of the contact structures caused a decrease in oxygen concentration in the Ga2O3 and an increase in the Ti. The annealing will cause formation of titanium oxide at the interface.49 There was also intermixing of the Ti and Au, with concomitant diffusion of Ga into the Au. It is difficult to form TiO and TiO2 by reaction between Ti and β-Ga2O3 since the formation enthalpy (ΔH°f) for Ga2O3 is −1089.1 kJ/mol,50 much lower than that of TiO and TiO2. However, Ti2O3 and Ti3O5 have very low ΔH°f, −1520.9 and −2459.4 kJ/mol, respectively, which are much lower than that of Ga2O3.50 After thermal annealing, Bae et al.49 showed a high atomic percentage of oxygen in the titanium layer (∼65%), consistent with formation of the Ti2O3 and Ti3O5 phases. In our case, the ITO interlayer should delay formation of these compounds to higher temperatures.

Fig. 4.

Sheet resistance, specific contact resistance and transfer resistance as a function of annealing temperature for Au/Ti/ITO/Ga2O3 contact stack.

Fig. 4.

Sheet resistance, specific contact resistance and transfer resistance as a function of annealing temperature for Au/Ti/ITO/Ga2O3 contact stack.

Close modal

The morphology degraded during annealing due to reactions of the metals, ITO and Ga2O3, as shown in the optical micrographs of Fig. 5. This shows contact stacks (a) as deposited, (b) annealed at 500 °C and (c) annealed at 600 °C. The annealing was carried out in N2 for 30 s at each temperature. All of these structures have the ITO etched off between pads and the images were taken using the same brightness and focus conditions on the microscope. When ITO is annealed on the doped Ga2O3 without any metallization above, no bubbling is observed in the ITO layer. The bubbling may be due in part to the ITO layer diffusing oxygen into the metal contact pads; however, similar bubbling was noted by Bae et al.49 in Ti/Au contacts on Ga2O3 flake, and was determined to be due to diffusion of oxygen into the contact. The ITO layer is only 10 nm, so likely a combined effect from the two layers causes the change in surface morphology. Using a diffusion barrier may prevent diffusion of gallium and oxygen atoms into the metal electrodes and prevent formation of TiO2. Bae et al.49 show transmission electron microscopy of pre- and postannealed contacts and depth dependent energy dispersive spectroscopy spectra, which show out-diffusion of gallium and oxygen atoms into the metallization. Optical microscopy images of their contacts as well note a bubbling, and degradation of the surface. Bae et al.49 observed the characteristics of contacts annealed in N2 and air and reported similar degradation of contact morphology after 700 °C annealing. Those contacts annealed in air do not demonstrate the same bubbling, as there is not a deficiency of oxygen atoms in the ambient during annealing.

Fig. 5.

(Color online) Optical micrographs of TLM patterns on Au/Ti/ITO/Ga2O3 (a) as-deposited, (b) after 500 °C anneal, and (c) 600 °C anneal. The pad separation is 10 μm in each case.

Fig. 5.

(Color online) Optical micrographs of TLM patterns on Au/Ti/ITO/Ga2O3 (a) as-deposited, (b) after 500 °C anneal, and (c) 600 °C anneal. The pad separation is 10 μm in each case.

Close modal

Figure 6 shows the effect of measurement temperature on the Au/Ti/ITO/Ga2O3 stack annealed at 500 °C in N2 for 30 s. The specific contact resistance decreased minimally with temperature, which is ideal for providing a consistent contact resistance in a device regardless of operating temperature. The thickness of 10 nm ITO was chosen based on the desire to reduce the conduction band difference between metal and Ga2O3. Thicker ITO would increase the vertical resistance, and thinner ITO may have pinholes to cover the Ga2O3. In our previous work, we identified rf-sputtered ITO/single crystal β-Ga2O3 heterostructures as having a Type I band alignment via x-ray photoelectron spectroscopy using the Kraut method.51,52 The individual component bandgaps were determined by reflection electron energy loss spectroscopy as 4.6 eV for Ga2O3 and 3.5 eV for ITO. The valence and conduction band offset was determined to be −0.78 ± 0.30 eV and −0.32 ± 0.13 eV, respectively. Shown in Fig. 7, the conduction band steps-down across the heterointerface favors electron transport and makes it easier to achieve an Ohmic contact. The Fermi level of the metal is hypothesized to have a larger offset on the conduction band of Ga2O3 than the ITO. Introducing an ITO interlayer allows for reducing the conduction band discontinuity between Ti and Ga2O3. Ohmic behavior was observed for Ti/Au deposited on ITO.53,54 The typical Schottky barrier heights of a variety of metallizations on Ga2O3 range from 0.9 to 1.45 eV.21,22,30,55–58 Use of a thin layer of ITO between a metal contact and the Ga2O3 is a promising approach for reducing contact resistance on Ga2O3-based power electronic devices and solar-blind photodetectors. Recent results on aluminum zinc oxide interlayers show similar results.59 Yao et al.60 concluded that metal work function was not a dominant factor in forming Ohmic contacts to Ga2O3, but that interfacial reactions were important.

Fig. 6.

Specific contact resistance, Rc, as a function of measurement temperature for Au/Ti/ITO/Ga2O3 contact stack initially annealed at 500 °C for 30 s.

Fig. 6.

Specific contact resistance, Rc, as a function of measurement temperature for Au/Ti/ITO/Ga2O3 contact stack initially annealed at 500 °C for 30 s.

Close modal
Fig. 7.

Schematic of band offset for Au/Ti/ITO on Ga2O3 and Au/Ti on Ga2O3 (Refs. 21, 22, 30, and 55–58).

Fig. 7.

Schematic of band offset for Au/Ti/ITO on Ga2O3 and Au/Ti on Ga2O3 (Refs. 21, 22, 30, and 55–58).

Close modal

The use of sputtered ITO interlayers in Ti/Au contacts to Ga2O3 improves electron transport across the heterointerface and enhances formation of an Ohmic contact. The band alignment of ITO on Ga2O3 is favorable for electron injection. For Au/Ti/ITO/Ga2O3 contact stacks, the minimum transfer resistance and specific contact resistance of 0.6 Ω mm and 6.3 × 10−5 Ω cm2 were achieved 600 °C annealing, respectively. Without the ITO, similar anneals did not lead to linear current-voltage characteristics of Au/Ti/Ga2O3 contact layers. In Ref. 27, high annealing temperatures of 900–1100 °C were used for forming Ohmic contacts—by using the ITO interlayer moderate annealing temperatures can be used. The use of the ITO interlayers is a convenient method for improving Ohmic contact resistance on n-type Ga2O3.

The project or effort depicted was also sponsored by the Department of the Defense, Defense Threat Reduction Agency, HDTRA1-17-1-011, monitored by Jacob Calkins. A portion of this research was conducted at the Center for Nanophase Materials Sciences, which is sponsored at Oak Ridge National Laboratory by the Office of Basic Energy Sciences, U.S. Department of Energy. The content of the information does not necessarily reflect the position or the policy of the federal government, and no official endorsement should be inferred. Part of the work at Tamura was supported by “The research and development project for innovation technique of energy conservation” of the New Energy and Industrial Technology Development Organization (NEDO), Japan. The authors also thank Kohei Sasaki from Tamura Corporation for fruitful discussions.

1.
H.
von Wenckstern
,
Adv. Electron. Mater.
3
,
1600350
(
2017
).
2.
S. I.
Stepanov
,
V. I.
Nikolaev
,
V. E.
Bougrov
, and
A. E.
Romanov
,
Rev. Adv. Mater. Sci.
44
,
63
(
2016
), available at http://www.ipme.ru/e-journals/RAMS/no_14416/06_14416_stepanov.pdf.
3.
M. A.
Mastro
,
A.
Kuramata
,
J.
Calkins
,
J.
Kim
,
F.
Ren
, and
S. J.
Pearton
,
ECS J. Solid State Sci. Technol.
6
,
P356
(
2017
).
4.
A.
Kuramata
,
K.
Koshi
,
S.
Watanabe
,
Y.
Yamaoka
,
T.
Masui
, and
S.
Yamakoshi
,
Jpn. J. Appl. Phys., Part 1
55
,
1202A2
(
2016
).
5.
M.
Higashiwaki
,
K.
Sasaki
,
H.
Murakami
,
Y.
Kumagai
,
A.
Koukitu
,
A.
Kuramata
,
T.
Masui
, and
S.
Yamakosh
,
Semicond. Sci. Technol.
31
,
034001
(
2016
).
6.
M.
Mohamed
,
C.
Janowitz
,
R.
Manzke
,
Z.
Galazka
,
R.
Uecker
,
R.
Fornari
,
J. R.
Weber
,
J. B.
Varley
, and
C. G.
Van de Walle
,
Appl. Phys. Lett.
97
,
211903
(
2010
).
7.
O.
Ueda
,
N.
Ikenaga
,
K.
Koshi
,
K.
Iizuka
,
A.
Kuramata
,
K.
Hanada
,
T.
Moribayashi
,
S.
Yamakoshi
, and
M.
Kasu
,
Jpn. J. Appl. Phys., Part 1
55
,
1202BD
(
2016
).
8.
M.
Higashiwaki
,
K.
Sasaki
,
A.
Kuramata
,
T.
Masui
, and
S.
Yamakoshi
,
Phys. Status Solidi A-
211
,
21
(
2014
).
9.
N. A.
Moser
,
J. P.
McCandless
,
A.
Crespo
,
K. D.
Leedy
,
A. J.
Green
,
E. R.
Heller
,
K. D.
Chabak
,
N.
Peixoto
, and
G. H.
Jessen
,
Appl. Phys. Lett.
110
,
143505
(
2017
).
10.
M.
Baldini
,
M.
Albrecht
,
A.
Fiedler
,
K.
Irmscher
,
R.
Schewski
, and
G.
Wagner
,
ECS J. Solid State Sci. Technol.
6
,
Q3040
(
2017
).
11.
A. J.
Green
 et al.,
IEEE Electron Device Lett.
37
,
902
(
2016
).
12.
M.
Tadjer
,
N. A.
Mahadik
,
V. D.
Wheeler
,
E. R.
Glaser
,
L.
Ruppalt
,
A. D.
Koehler
,
K. D.
Hobart
,
C. R.
Eddy
, Jr.
, and
F. J.
Kub
,
ECS J. Solid State Sci. Technol.
5
,
P468
(
2016
).
13.
M. H.
Wong
,
K.
Sasaki
,
A.
Kuramata
,
S.
Yamakoshi
, and
M.
Higashiwaki
,
IEEE Electron Device Lett.
37
,
212
(
2016
).
14.
K.
Zheng
,
J. S.
Wallace
,
C.
Hemiburger
,
K.
Sasaki
,
A.
Kuramata
,
T.
Masui
,
J. S.
Gardella
, and
U.
Singisetti
,
IEEE Electron Device Lett.
38
,
513
(
2017
).
15.
S.
Rafique
,
L.
Han
,
M. J.
Tadjer
,
J. A.
Freitas
, Jr.
,
N. A.
Mahadik
, and
H.
Zhao
,
Appl. Phys. Lett.
108
,
182105
(
2016
).
16.
M. H.
Wong
,
K.
Sasaki
,
A.
Kuramata
,
S.
Yamakoshi
, and
M.
Higashiwaki
,
Appl. Phys. Lett.
106
,
032105
(
2015
).
17.
S.
Oh
,
G.
Yang
, and
J.
Kim
,
ECS J. Solid State Sci. Technol.
6
,
Q3022
(
2017
).
18.
W. S.
Hwang
 et al.,
Appl. Phys. Lett.
104
,
203111
(
2014
).
19.
S.
Ahn
,
F.
Ren
,
J.
Kim
,
S.
Oh
,
J.
Kim
,
M. A.
Mastro
, and
S. J.
Pearton
,
Appl. Phys. Lett.
109
,
062102
(
2016
).
20.
J.
Kim
,
S.
Oh
,
M.
Mastro
, and
J.
Kim
,
Phys. Chem. Chem. Phys.
18
,
15760
(
2016
).
21.
S.
Oh
,
Y.
Jung
,
M. A.
Mastro
,
J. K.
Hite
,
C. R.
Eddy
, Jr.
, and
J.
Kim
,
Opt. Express
23
,
28300
(
2015
).
22.
A. M.
Armstrong
,
M. H.
Crawford
,
A.
Jayawardena
,
A.
Ahyi
, and
S.
Dhar
,
J. Appl. Phys.
119
,
103102
(
2016
).
23.
T. J.
Anderson
,
J. D.
Greenlee
,
B. N.
Feigelson
,
J. K.
Hite
,
F. J.
Kub
, and
K. D.
Hobart
,
ECS J. Solid State Sci. Technol.
5
,
Q176
(
2016
).
24.
J.
Anaya
 et al.,
Acta Mater.
103
,
141
(
2016
).
25.
D. J.
Meyer
 et al.,
IEEE Electron Device Lett.
35
,
1013
(
2014
).
26.
E. A.
Imhoff
,
F. J.
Kub
,
K. D.
Hobart
,
M. G.
Ancona
,
B. L.
VanMil
,
D. K.
Gaskill
,
K.
Lew
,
R. L.
Myers-Ward
, and
C. R.
Eddy
,
IEEE Trans. Electron Devices
58
,
3395
(
2011
).
27.
K.
Sasaki
,
M.
Higashiwaki
,
A.
Kuramata
,
T.
Masui
, and
S.
Yamakoshi
,
IEEE Electron. Dev. Lett.
34
,
493
(
2013
).
28.
T.
Oshima
 et al.,
Jpn. J. Appl. Phys., Part 1
55
,
1202B7
(
2016
).
29.
E. G.
Víllora
,
K.
Shimamura
,
Y.
Yoshikawa
,
T.
Ujiie
, and
K.
Aoki
,
Appl. Phys. Lett.
92
,
202120
(
2008
).
30.
M.
Mohamed
,
K.
Irmscher
,
C.
Janowitz
,
Z.
Galazka
,
R.
Manzke
, and
R.
Fornari
,
Appl. Phys. Lett.
101
,
132106
(
2012
).
31.
V. D.
Wheeler
,
D. I.
Shahin
,
M. J.
Tadjer
, and
C. R.
Eddy
, Jr.
,
ECS J. Solid State Sci. Technol.
6
,
Q3052
(
2017
).
32.
Z.
Che
,
K.
Nishhagi
,
X.
Wang
,
K.
Saito
,
T.
Tanaka
,
M.
Nishino
,
M.
Arita
, and
Q.
Guo
,
Appl. Phys. Lett.
109
,
102106
(
2016
).
33.
T. C.
Lovejoy
 et al.,
Appl. Phys. Lett.
100
,
181602
(
2012
).
34.
A.
Navarro-Quezada
,
Z.
Galazka
,
S.
Alamé
,
D.
Skuridina
,
P.
Vogt
, and
N.
Esser
,
Appl. Surf. Sci.
349
,
368
(
2015
).
35.
S.
Muller
,
H.
von Wenckstern
,
F.
Schmidt
,
D.
Splith
,
H.
Frenzel
, and
M.
Grundmann
,
Semiconduct. Sci. Technol.
32
,
065013
(
2017
).
36.
S.
Müller
,
H.
von Wenckstern
,
F.
Schmidt
,
D.
Splith
,
F.-L.
Schein
,
H.
Frenzel
, and
M.
Grundmann
,
Appl. Phys. Express
8
,
121101
(
2015
).
37.
S.
Ahn
,
F.
Ren
,
L.
Yuan
,
S. J.
Pearton
, and
A.
Kuramata
,
ECS J. Solid State Sci. Technol.
6
,
P68
(
2017
).
38.
D.
Guo
 et al.,
Opt. Mater. Express
4
,
1067
(
2014
).
39.
D. Y.
Guo
,
Z. P.
Wu
,
Y. H.
An
,
X. C.
Guo
,
X. L.
Chu
,
C. L.
Sun
,
L. H.
Li
,
P. G.
Li
, and
W. H.
Tangless
,
Appl. Phys. Lett.
105
,
023507
(
2014
).
40.
S. J.
Kim
,
IEEE Photonics Technol. Lett.
17
,
1617
(
2005
).
41.
P. D. C.
King
and
T. D.
Veal
,
J. Phys. C: Condens. Matter
23
,
334214
(
2011
).
42.
T.
Minami
and
T.
Miyata
,
Thin Solid Films
517
,
1474
(
2008
).
43.
G. J.
Exarhos
and
X.-D.
Zhou
,
Thin Solid Films
515
,
7025
(
2007
).
44.
H.
Liu
,
V.
Avrutin
,
N.
Izyumskaya
,
Ü.
Özgür
, and
H.
Morkoç
,
Superlattice Microstruct.
48
,
458
(
2010
).
45.
T.
Minami
,
Oxide Semiconductors
, Semiconductors and Semimetals Vol.
88
, edited by
B. G.
Svensson
,
S. J.
Pearton
, and
C.
Jagadish
(
Academic
,
San Diego
,
2013
), Chap. 5.
46.
D.
Ginley
,
H.
Hosono
, and
D. C.
Paine
,
Handbook of Transparent Conductors
(
Springer
,
Berlin, Germany
,
2011
).
47.
A.
Suzuki
,
T.
Matsushita
,
T.
Aoki
, and
Y.
Yoneyama
,
Jpn. J. Appl. Phys., Part 2
40
,
L410
(
2001
).
48.
D. K.
Schroder
,
Semiconductor Material and Device Characterization
, 3rd ed. (
Wiley
,
Hoboken, NJ
,
2006
).
49.
J.
Bae
,
H.-Y.
Kim
, and
J.
Kim
,
ECS J. Solid State Sci. Technol.
6
,
Q3045
(
2017
).
50.
W. M.
Haynes
,
CRC Handbook of Chemistry and Physics
, 93rd ed. (
CRC
,
Boca Raton
,
2016
).
51.
P. H.
Carey
, IV
,
F.
Ren
,
D. C.
Hays
,
B. P.
Gila
,
S. J.
Pearton
,
S.
Jang
, and
A.
Kuramata
,
Appl. Surf. Sci.
422
,
179
(
2017
).
52.
E. A.
Kraut
,
R. W.
Grant
,
J. R.
Waldrop
, and
S. P.
Kowalczyk
,
Phys. Rev. Lett.
44
,
1620
(
1980
).
53.
B.
Kang
,
J.
Chen
,
F.
Ren
,
Y.
Li
,
H.
Kim
,
D.
Norton
, and
S. J.
Pearton
,
Appl. Phys. Lett.
88
,
182101
(
2006
).
54.
M.
Kim
and
T.
Kim
,
ACS Appl. Mater. Interfaces
8
,
5453
(
2016
).
55.
D.
Splith
,
S.
Müller
,
F.
Schmidt
,
H.
von Wenckstern
,
J. J.
van Rensburg
,
W. E.
Meyer
, and
M.
Grundmann
,
Phys. Status Solidi A
211
,
40
(
2014
).
56.
Z.
Zhang
,
E.
Farzana
,
A. R.
Arehart
, and
S. A.
Ringel
,
Appl. Phys. Lett.
108
,
052105
(
2016
).
57.
S.
Krishnamoorthy
 et al.,
Appl. Phys. Lett.
111
,
023502
(
2017
).
58.
K.
Sasaki
,
M.
Higashiwaki
,
A.
Kuramata
,
T.
Masui
, and
S.
Yamakoshi
,
Appl. Phys. Express
6
,
086502
(
2013
).
59.
P. H.
Carey
,
J.
Yang
,
F.
Ren
,
D. C.
Hays
,
S. J.
Pearton
,
S.
Jang
,
A.
Kuramata
, and
I. I.
Kravchenko
,
AIP Adv.
7
,
095313
(
2017
).
60.
Y.
Yao
,
R. F.
Davis
, and
L. M.
Porter
,
J. Electron. Mater.
46
,
2053
(
2017
).