The Schottky barriers of Ti, Mo, Co, Ni, Pd, and Au on (100) β-Ga2O3 substrates were analyzed using a combination of current-voltage (J-V), capacitance-voltage (C-V), and current-voltage-temperature (J-V-T) measurements. Near-ideal, average ideality factors for Ti, Mo, Co, and Ni were 1.05–1.15, whereas higher ideality factors (∼1.3) were observed for Pd and Au contacts. Barrier heights ranging from 0.60 to 1.20 eV were calculated from J-V measurements for the metals with low ideality factors. C-V measurements of all Schottky metals were conducted and yielded average barrier heights ranging from 0.78 to 1.98 eV. J-V-T measurements of Ti and Co diodes yielded barrier heights of 0.81 and 1.35 eV, respectively. The results reveal a strong positive correlation between the calculated Schottky barrier heights and the metal work functions: the index of interface behavior, S = 0.70, 0.97, and 0.81 for J-V, C-V, and J-V-T data, respectively.

Beta-gallium oxide (β-Ga2O3) is a semiconductor that has garnered interest over the past decade due to its ultrawide bandgap (∼4.6–4.9 eV),1 wide range of n-type doping2–5 (Si, Ge, and Sn), and its ability to be produced as large-area, melt-grown substrates that serve as a platform for device-quality epitaxial layers.6–8 For these reasons, Ga2O3 is being pursued for high power devices9 and UV photodetectors.10 Other wide bandgap semiconductors such as GaN and SiC suffer from higher costs to produce substrates from vapor phase growth methods. In addition, Ga2O3 has a higher theoretical breakdown field (∼8 MV/cm) and Baliga figure of merit (∼3400) than GaN and SiC.11 

Studies of Schottky contacts to β-Ga2O3 play a fundamental role in designing metal-semiconductor field-effect transistors12 and Schottky diodes.13 As such, it is important to develop contacts with high Schottky barrier heights and low leakage currents. Metals with a wide range of work functions have been explored as Schottky contacts to Ga2O3, including W,14 Cu,15 Ni,16 Au,17 Pt,18 TiN,19 Pd,20 Ir,21 Ag,20 and Mo.22 Similarly, Schottky contacts to (100), (010), (001), and (2¯01) orientations of β-Ga2O3 substrates grown via edge-defined film fed growth,23,24 Czochralski,16,25 and floating zone26 and to β-Ga2O3 epitaxial layers grown via metalorganic chemical vapor deposition,27–30 halide vapor phase epitaxy,6,28 and pulsed laser deposition31 have been reported. Yao et al. investigated five different Schottky metals to (2¯01) β-Ga2O3 substrates and found that the metal work function was not a strong indicator of the Schottky barrier height.18 In a subsequent study, authors examined oxidized and pure (unoxidized) metal contacts to (2¯01) β-Ga2O3 and also concluded that the work function was not a strong indicator of Schottky barrier height,20 whereas Farzana et al. established evidence that the Fermi level is not completely pinned for Ni, Pd, and Pt Schottky contacts to the (010) β-Ga2O3 surface,32 and other data suggest that the metal work function is not a strong indicator of Schottky barrier height on the (010) β-Ga2O3 surface.21 There is limited published work on Schottky contacts on the (100)16,17,33–35 and (001)36,37 orientations of β-Ga2O3. In this work, we measured the electrical behavior, at both room temperature and as a function of elevated temperature, of Ti, Mo, Co, Ni, Pd, and Au Schottky contacts to (100) β-Ga2O3 and calculated the Schottky barrier heights and ideality factors from these measurements. Observed trends are discussed and compared with previously published results for other Ga2O3 surfaces.

N-type, unintentionally doped (100) β-Ga2O3 single crystal wafers were grown by the Czochralski method at the Leibniz-Institut für Kristallzüchtung (IKZ) in Berlin, Germany. The as-received samples were 1 cm2 and had resistivity, mobility, and free-carrier concentration of 0.2 Ω cm, 60 cm2/V s, and 5 × 1017 cm−3, respectively, as determined from Hall effect measurements. The doping concentration calculated from the C-V measurements of Schottky diodes was 2.5–4.5 × 1017 cm−3.

All metals were deposited onto unheated substrates by electron-beam evaporation (base pressure of ∼10−8 Torr). Ohmic contacts were formed using Ti/Au (20/100 nm) annealed at 450 °C, either for 1 min in N2 in a rapid thermal annealing furnace (for backside ohmic contacts, referred to below as vertical structures) or for 5 min at 400 °C in Ar in a resistively heated quartz tube furnace (for frontside ohmic contacts, referred to as the lateral structures). Prior to deposition of all Schottky contacts, each substrate was soaked in 10% hydrochloric acid (HCl) for 5 min, rinsed in de-ionized (DI) water, and immersed for 5 min in boiling H2O2 at 85 °C. All samples were blown dry in nitrogen after each DI water rinse. The Schottky metals were subsequently deposited through a Mo shadow mask to form circular Schottky contacts with diameters of 125, 250, and 500 μm. Ti, Co, and Pd Schottky contacts were fabricated as vertical device structures with backside ohmic contacts, whereas Mo, Ni, and Au were fabricated as lateral structures with frontside ohmic contacts. The thickness of all deposited Schottky contacts is 30 nm. The Ti Schottky contacts were coated with 50 nm of Au to serve as a passivation layer.

Hall measurements were acquired using an MMR Technologies Hall Measurement System at room temperature. For these measurements, Ti/Au (20/50 nm) contacts were deposited in a van der Pauw pattern onto a (100) β-Ga2O3 substrate and annealed at 400 °C for 5 min in Ar. J-V and J-V-T measurements were obtained using an Agilent 4155C Semiconductor Parameter Analyzer and a Signatone S-1160A-4N probe station. C-V measurements were performed at 1 MHz using an HP 4284 LCR meter and a Signatone S-1160A-4N probe station.

For Schottky contacts that follow the thermionic emission model, the current density (J) vs voltage (V) behavior is described in the following equation:

J(V)=Js[eq(VIRs)nkT1],
(1)

where q is the electronic charge, I is the current, RS is the series resistance, n is the ideality factor, and k is the Boltzmann constant. Js is the saturation current density defined as follows:

Js=AT2eqϕBJVkT,
(2)

where A** is Richardson's constant for the semiconductor and ϕBJV is the J-V determined Schottky barrier height. The Richardson constant for β-Ga2O3 has been calculated to be 33.65 A/cm2 K2.38 

Figure 1(a) shows representative log-J versus V characteristics for each Schottky metal under forward bias conditions. Most metals display linearity over several decades of current before reaching series resistance limitations. An exception is Au, which displays multiple linear regions, suggestive of an inhomogeneous Schottky barrier, as established by other authors on other β-Ga2O3 surfaces.18,39

FIG. 1.

(a) J-V and (b) C-V characteristics for Ti, Mo, Co, Ni, Pd, and Au Schottky contacts on (100) β-Ga2O3. The C-V data are plotted as diode area-squared divided by capacitance-squared vs voltage to allow calculation of Schottky barrier heights for each metal. The diameter of each diode is 500 μm.

FIG. 1.

(a) J-V and (b) C-V characteristics for Ti, Mo, Co, Ni, Pd, and Au Schottky contacts on (100) β-Ga2O3. The C-V data are plotted as diode area-squared divided by capacitance-squared vs voltage to allow calculation of Schottky barrier heights for each metal. The diameter of each diode is 500 μm.

Close modal

Ideality factors and Schottky barrier heights were calculated from J-V characteristics using the Cheung and Cheung method.40 This method uses a plot of

dVdlnJ=JARSnkTq,
(3)

where A = diode area to extract n from the y-intercept of a dVdlnJ versus J plot; and a function

H(J)VnkTqln(JAT2)=JARSnϕBJV
(4)

from which ϕBJV is determined from the y-intercept of a H(J) versus J plot. Table I lists the ϕBJV values for Ti, Mo, Co, and Ni Schottky metal contacts, along with the ideality factors and the leakage current density values at −5 V for all Schottky contacts. Because of the higher ideality factors for Pd and Au contacts, the calculated ΦBJV values (1.42 ± 0.07 and 1.37 ± 0.09 eV, respectively) for these contacts do not follow thermionic emission and are not included in the analysis. The higher ideality factors for the Au and Pd contacts are associated with spatial inhomogeneity of the Schottky barriers for Au/Ga2O3 and Pd/Ga2O3 interfaces. Spatially inhomogeneous Schottky barriers are believed to result from Au-Ga and Pd-Ga alloy formation at the metal semiconductor interface. Pd-Ga and Au-Ga elemental formations have been reported on other Ga-based semiconductors including GaAs41,42 and GaN.43,44 The larger ideality factors observed for the Pd and Au contacts would be compatible with this phenomenon. Others have also reported evidence for an inhomogeneous Schottky barrier in Au Schottky contacts to (010) β-Ga2O3.32 Furthermore, our analysis of J-V-T measurements of Pd contacts, described below and in the supplementary material,51 indicates the presence of significant barrier inhomogeneity, as quantified by the standard deviation of the gaussian distribution, σ0. The σ0 value of the barrier shows a higher value for Pd than both Co and Ti, implying more variance in the measurements of the Schottky barrier height for Pd.

TABLE I.

Electrical properties extracted from J-V, C-V, and J-V-T measurements. Where applicable, the standard deviations are indicated with ± values.

MetalLeakage current @−5V(A/cm2)nϕBJV(eV)ϕBHom(eV)ϕBCV(eV)ϕBJVT(eV)
Ti 7.64 × 10−2 1.15 ± 0.05 0.60 ± 0.03 0.68 0.78 ± 0.08 0.81 ± 0.11 
Mo 1.05 × 10−5 1.05 ± 0.02 1.00 ± 0.02 0.99 0.95 ± 0.05 — 
Co 6.67 × 10−9 1.06 ± 0.01 1.20 ± 0.02 1.32 1.33 ± 0.10 1.35 ± 0.03 
Ni 3.32 × 10−8 1.05 ± 0.02 1.20 ± 0.03 1.25 1.25 ± 0.05 — 
Pd 1.59 × 10−9 1.26 ± 0.07 a 1.62 1.86 ± 0.13 a 
Au 4.77 × 10−9 1.32 ± 0.07 a 1.41 1.98 ± 0.04 — 
MetalLeakage current @−5V(A/cm2)nϕBJV(eV)ϕBHom(eV)ϕBCV(eV)ϕBJVT(eV)
Ti 7.64 × 10−2 1.15 ± 0.05 0.60 ± 0.03 0.68 0.78 ± 0.08 0.81 ± 0.11 
Mo 1.05 × 10−5 1.05 ± 0.02 1.00 ± 0.02 0.99 0.95 ± 0.05 — 
Co 6.67 × 10−9 1.06 ± 0.01 1.20 ± 0.02 1.32 1.33 ± 0.10 1.35 ± 0.03 
Ni 3.32 × 10−8 1.05 ± 0.02 1.20 ± 0.03 1.25 1.25 ± 0.05 — 
Pd 1.59 × 10−9 1.26 ± 0.07 a 1.62 1.86 ± 0.13 a 
Au 4.77 × 10−9 1.32 ± 0.07 a 1.41 1.98 ± 0.04 — 
a

The ideality factors obtained from I-V measurements were too large to attribute ϕBJV to pure thermionic emission transport.

Schottky barrier heights were also calculated from C-V measurements. As derived from depletion approximation,45 

A2C(V)2=2(qϕBCVkTln(NCNd)qVkT)q2εsNd,
(5)

where ϕBCV, εs, and Nd are the C-V determined Schottky barrier height, semiconductor permittivity, and doping concentration, respectively. Nc, the conduction band density of states, is expressed as follows:

Nc=2(2πmkTh2)32,
(6)

where m=0.28m0 is the electron effective mass in β-Ga2O3 and m0 is the free electron mass. For β-Ga2O3 at room temperature, the conduction band density of states is calculated to be Nc=2.67x1018cm3. Therefore, by plotting A2/C2 versus V [as shown in Fig. 1(b)], Nd can be calculated from the slope and the barrier height can be calculated from the x-intercept.

The C-V determined Schottky barrier heights are listed in Table I, column 6. Note that differences in C-V and J-V determined barrier heights are commonly reported. For example, barrier heights from C-V measurements are typically higher than those calculated from J-V measurements. Tung46 attributes this effect to slight nonidealities (or spatially inhomogeneous Schottky barriers) in the diodes. For Mo, Co, and Ni, the agreement between the C-V and J-V determined barrier heights are within 3%–9%, whereas the agreement between these values for Ti is ∼20%–30%.

A homogeneous Schottky barrier height, ϕBHom, can be defined by plotting ϕBJV versus n and extrapolating to n = 1, as shown in Fig. 2. This method47 works well for diodes with low ideality factors. For example, the ϕBHom values (Table I, column 5) for Ti, Mo, Co, and Ni are in relatively close agreement with ϕBJV and ϕBCV values. For Pd, there is also a reasonably close agreement between ϕBHomand ΦBCV, but for Au, these values differ by ∼0.5 eV.

FIG. 2.

J-V determined barrier heights for Ti, Mo, Co, Ni, Pd, and Au on (100) β-Ga2O3 plotted vs ideality factor.

FIG. 2.

J-V determined barrier heights for Ti, Mo, Co, Ni, Pd, and Au on (100) β-Ga2O3 plotted vs ideality factor.

Close modal

We accounted for spatial distribution of Schottky barriers in our analysis of J-V-T measurements of the vertical (Ti, Co, and Pd) device structures. As such, J-V-T measurements provided a potential third method to determine Schottky barrier heights for these diodes. However, as evidenced by the high ideality factors of the Pd diodes (∼1.3), the current transport is not dominated by thermionic emission. Although minimum ideality factors of 1.10–1.15 are obtained in the 150–200 °C range was observed, corresponding to a calculated barrier height of ∼1.7 eV at that temperature, the generally high ideality factors precluded a determination of a reliable ΦBJVT for the Pd contacts.

The J-V-T data for the vertical devices were fit using a model developed by Werner and Güttler,48 which assumes the measured Schottky barriers yield a normalized Gaussian spatial distribution, P(ϕB), with average barrier height, ϕB¯, and standard deviation, σ,

P(ϕB)=1σ2πe(ϕB¯ϕB)22σ2.
(7)

By applying this model as detailed in the supplementary materials,51 we obtain the following J-V-T relationship:

ln(JsT2)q2σ022k2T2=lnAqϕBJVTkT,
(8)

where σ0 is the standard deviation at zero bias and serves as a metric that determines the degree of spatial inhomogeneity of the Schottky barrier and also represents how strongly the Schottky barrier changes with temperature. σ0 was determined to be 92, 72, and 197 meV for the Ti, Co, and Pd contacts, respectively. Figure 3 plots the left-hand side of Eq. (8) versus 1 /kT, which provides a graphical method to calculate both Richardson's constant of β-Ga2O3 (see Fig. 3 inset table) and the ϕBJVTvalues (listed in the last column of Table I). The Richardson's constant values calculated from this experimental data are in reasonable agreement with the theoretical value of β-Ga2O3 of 33.65 A/cm2 K2.38 The near linear behavior of the curves in Fig. 3 indicates that this model fits our data well.

FIG. 3

(a) J-V-T characteristics for Schottky diodes with Ti and Co on (100) β-Ga2O3 plotted in accordance with the inhomogeneous Schottky barrier model described in Eq. (8).

FIG. 3

(a) J-V-T characteristics for Schottky diodes with Ti and Co on (100) β-Ga2O3 plotted in accordance with the inhomogeneous Schottky barrier model described in Eq. (8).

Close modal

The Schottky barrier heights calculated from the J-V, C-V, and J-V-T measurements discussed above are plotted versus the metal work functions in Fig. 4. The black dashed line in the lower part of the plot represents the Schottky–Mott theory for an ideal metal-semiconductor contact,

ϕB=ϕMχs,
(9)

where ΦM is the work function of the metal and χS is the electron affinity of the semiconductor; in this case, χS for β-Ga2O3 was assumed to be 4.05 eV.38 

FIG. 4.

Schottky barrier heights of Ti, Mo, Co, Ni, Pd, and Au on (100) β-Ga2O3 plotted vs metal work function. The black dashed line represents the Schottky–Mott equation using an electron affinity value of 4.05 eV.

FIG. 4.

Schottky barrier heights of Ti, Mo, Co, Ni, Pd, and Au on (100) β-Ga2O3 plotted vs metal work function. The black dashed line represents the Schottky–Mott equation using an electron affinity value of 4.05 eV.

Close modal

According to the seminal study by Kurtin et al.,49 Ga2O3 is expected to closely follow this predicted Schottky–Mott behavior due to its high degree of ionicity; that is, the slope of the plot, S=dϕBdϕM, termed the index of interface behavior, should be close to 1.00. However, much research since that time has shown other factors can play a dominant role: for example, large variations in S values for different surface planes of the same semiconductor, e.g., SiC and Ga2O3, or for different forms of the same semiconductor, e.g., SnS and diamond/NCD.50 A least squares fit to each set of data produced the following lines:

ϕBJV=0.70ϕM2.31eV,R2=0.89,
(10)
ϕBCV=0.97ϕM3.50eV,R2=0.89.
(11)

The corresponding S values are SJV = 0.70 and SCV = 0.97. Furthermore, for the two data points from J-V-T measurements, SJVT = 0.82.

As discussed earlier, differences between Schottky barrier heights (and therefore, S values) calculated from different measurement methods can be attributed to the presence of Schottky barrier inhomogeneities, particularly for the contacts (Pd and Au) with higher ideality factors. Overall, however, the results from this study on (100) β-Ga2O3 reveal a strong positive correlation between the calculated Schottky barrier heights and the work function of the metal contacts. In contrast, previous studies of metal contacts to (2¯01) β-Ga2O3 did not demonstrate a discernable correlation between Schottky barrier height and metal work function,18,20 whereas results on (010) β-Ga2O3 have been mixed.21,32 Hou et al.20 attribute the strong Fermi level pinning on the (2¯01) Ga2O3 surface to its higher oxygen dangling bond density and oxygen vacancies. Our conclusion from the present study in comparison with previous studies discussed above is that the surface plane/orientation of β-Ga2O3 has a significant effect on the electrical properties (e.g., Schottky barrier height) of metal contacts to this ultrawide bandgap semiconductor.

In summary, Ti, Mo, Co, Ni, Pd, and Au Schottky contacts to (100) Czochralski-grown β-Ga2O3 were electrically characterized via J-V, C-V, and/or J-V-T measurements. Analysis of the measurements revealed a strong correlation between the calculated Schottky barrier heights and the metal work functions. The electrical behavior of the Au and Pd contacts are associated with the presence of spatially inhomogeneous Schottky barriers, which require further investigation.

This material is based upon work supported by the Air Force Office of Scientific Research (Program Manager, Ali Sayer) under Award No. FA9550-18-1-0387. This work was partially performed in the framework of GraFOx a Leibniz-Science Campus partially funded by the Leibniz Association.

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

1.
C.
Janowitz
 et al,
New J. Phys.
13
,
085014
(
2011
).
2.
Z.
Galazka
 et al,
J. Cryst. Growth
486
,
82
(
2018
).
3.
Z.
Galazka
 et al,
J. Cryst. Growth
404
,
184
(
2014
).
4.
H.
Aida
,
K.
Nishiguchi
,
H.
Takeda
,
N.
Aota
,
K.
Sunakawa
, and
Y.
Yaguchi
,
Jpn. J. Appl. Phys.
47
,
8506
(
2008
).
5.
Z.
Galazka
 et al,
J. Cryst. Growth
529
,
125297
(
2020
).
6.
H.
Murakami
 et al,
Appl. Phys. Express
8
,
015503
(
2015
).
7.
Z.
Galazka
,
Semicond. Sci. Technol.
33
,
113001
(
2018
).
8.
J. D.
Blevins
,
K.
Stevens
,
A.
Lindsey
,
G.
Foundos
, and
L.
Sande
,
IEEE Trans. Semicond. Manuf.
32
,
466
472
(
2019
).
9.
K.
Konishi
,
K.
Goto
,
H.
Murakami
,
Y.
Kumagai
,
A.
Kuramata
,
S.
Yamakoshi
, and
M.
Higashiwaki
,
Appl. Phys. Lett.
110
,
103506
(
2017
).
10.
L. A. M.
Lyle
,
S.
Okur
,
V. S. N.
Chava
,
M. L.
Kelley
,
R. F.
Davis
,
G. S.
Tompa
,
M. V. S.
Chandrashekhar
,
A. B.
Greytak
, and
L. M.
Porter
,
J. Electron. Mater.
49
(
6
),
3490
3498
(
2020
).
11.
L. A. M.
Lyle
,
Research and Development of Electrical Contacts to β-Ga2O3 for Power Electronics and UV Photodetectors
(
Carnegie Mellon University
,
2020
).
12.
Z.
Liu
,
P.-G.
Li
,
Y.-S.
Zhi
,
X.-L.
Wang
,
X.-L.
Chu
, and
W.-H.
Tang
,
Chin. Phys. B
28
,
1
17
(
2019
).
13.
L. A. M.
Lyle
,
L.
Jiang
,
K. K.
Das
, and
L. M.
Porter
, in
Gallium Oxide—Technology, Devices and Applications
, edited by
S. J.
Pearton
,
F.
Ren
, and
M. A.
Mastro
(
Elsevier
,
Cambridge, MA
,
2019
), p.
231
.
14.
C.
Fares
,
F.
Ren
, and
S. J.
Pearton
,
ECS J. Solid State Sci. Technol.
8
,
Q3007
(
2018
).
15.
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
).
16.
K.
Irmscher
,
Z.
Galazka
,
M.
Pietsch
,
R.
Uecker
, and
R.
Fornari
,
J. Appl. Phys.
110
,
063720
(
2011
).
17.
M.
Mohamed
,
K.
Irmscher
,
C.
Janowitz
,
Z.
Galazka
,
R.
Manzke
, and
R.
Fornari
,
Appl. Phys. Lett.
101
,
132106
(
2012
).
18.
Y.
Yao
,
R.
Gangireddy
,
J.
Kim
,
K. K.
Das
,
R. F.
Davis
, and
L. M.
Porter
,
J. Vac. Sci. Technol. B
35
,
03D113
(
2017
).
19.
M. J.
Tadjer
,
V. D.
Wheeler
,
D. I.
Shahin
,
C. R.
Eddy
, and
F. J.
Kub
,
ECS J. Solid State Sci. Technol.
6
,
P165
(
2017
).
20.
C.
Hou
,
R. M.
Gazoni
,
R. J.
Reeves
, and
M. W.
Allen
,
Appl. Phys. Lett.
114
,
033502
(
2019
).
21.
C.
Hou
,
R. M.
Gazoni
,
R. J.
Reeves
, and
M. W.
Allen
,
IEEE Electron Device Lett.
40
,
337
(
2019
).
22.
Z.
Hu
 et al,
Nanoscale Res. Lett.
14
,
2
(
2019
).
23.
A.
Kuramata
,
K.
Koshi
,
S.
Watanabe
,
Y.
Yamaoka
,
T.
Masui
, and
S.
Yamakoshi
,
Jpn. J. Appl. Phys.
55
,
1202A2
(
2016
).
24.
K.
Shimamura
,
E. G.
Víllora
,
T.
Ujiie
, and
K.
Aoki
,
Appl. Phys. Lett.
92
,
201914
(
2008
).
25.
M.
Mohamed
 et al,
Appl. Phys. Lett.
97
,
211903
(
2010
).
26.
N.
Suzuki
,
S.
Ohira
,
M.
Tanaka
,
T.
Sugawara
,
K.
Nakajima
, and
T.
Shishido
,
Phys. Status Solidi C
4
,
2310
(
2007
).
27.
Y.
Yao
,
L. A. M.
Lyle
,
J. A.
Rokholt
,
S.
Okur
,
G. S.
Tompa
,
T.
Salagaj
,
N.
Sbrockey
,
R. F.
Davis
, and
L. M.
Porter
,
ECS Trans.
80
,
191
(
2017
).
28.
Y.
Yao
,
S.
Okur
,
L. A. M.
Lyle
,
G. S.
Tompa
,
T.
Salagaj
,
N.
Sbrockey
,
R. F.
Davis
, and
L. M.
Porter
,
Mater. Res. Lett.
6
,
268
(
2018
).
29.
S.
Bin Anooz
 et al,
Appl. Phys. Lett.
116
,
182106
(
2020
).
30.
S.
Bin Anooz
 et al,
J. Phys. D
54
,
034003
(
2021
).
31.
K. D.
Leedy
 et al,
Appl. Phys. Lett.
111
,
012103
(
2017
).
32.
E.
Farzana
,
Z.
Zhang
,
P. K.
Paul
,
A. R.
Arehart
, and
S. A.
Ringel
,
Appl. Phys. Lett.
110
,
202102
(
2017
).
33.
K.
Jiang
 et al,
ECS Trans.
92
,
71
(
2019
).
34.
L.
Du
 et al,
IEEE Electron Device Lett.
40
,
451
(
2019
).
35.
G.
Jian
 et al,
AIP Adv.
8
,
015316
(
2018
).
36.
M.
Higashiwaki
,
K.
Sasaki
,
A.
Kuramata
,
T.
Masui
, and
S.
Yamakoshi
,
Appl. Phys. Lett.
100
,
013504
(
2012
).
37.
M.
Higashiwaki
 et al,
Appl. Phys. Lett.
108
,
133503
(
2016
).
38.
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
).
39.
E.
Farzana
,
E.
Ahmadi
,
J. S.
Speck
,
A. R.
Arehart
, and
S. A.
Ringel
,
J. Appl. Phys.
123
,
161410
(
2018
).
40.
S. K.
Cheung
and
N. W.
Cheung
,
Appl. Phys. Lett.
49
,
85
(
1986
).
41.
H.-Y.
Nie
and
Y.
Nannichi
,
Jpn. J. Appl. Phys.
30
,
906
(
1991
).
42.
Z.
Liliental-Weber
,
J.
Washburn
,
N.
Newman
,
W. E.
Spicer
, and
E. R.
Weber
,
Appl. Phys. Lett.
49
,
1514
(
1986
).
43.
H.
Jung
,
R.
Behtash
,
J. R.
Thorpe
,
K.
Riepe
,
F.
Bourgeois
,
H.
Blanck
,
A.
Chuvilin
, and
U.
Kaiser
,
Phys. Status Solidi C
6
,
S976
(
2009
).
44.
T.
Jang
,
S. N.
Lee
,
O. H.
Nam
, and
Y.
Park
,
Appl Phys. Lett.
88
,
193505
(
2006
).
45.
S. M.
Sze
and
K. N.
Kwok
,
Phys. Semicond. Devices
763
,
170
181
(
2007
).
46.
R. T.
Tung
,
Phys. Rev. B: Condens. Matter
45
,
13509
(
1992
).
47.
R. F.
Schmitsdorf
,
J. Vac. Sci. Technol. B
15
,
1221
(
1997
).
48.
J. H.
Werner
and
H. H.
Güttler
,
J. Appl. Phys.
69
,
1522
(
1991
).
49.
S.
Kurtin
,
T. C.
McGill
, and
C. A.
Mead
,
Phys. Rev. Lett.
22
,
1433
(
1969
).
50.
L. M.
Porter
and
J. R.
Hajzus
,
J. Vac. Sci. Technol. A
38
,
031005
(
2020
).
51.
See supplementary material detailing the J-V-T analysis at http://dx.doi.org/10.1116/6.0000877.

Supplementary Material