Ultrawide bandgap β-(AlxGa1−x)2O3 vertical Schottky barrier diodes on (010) β-Ga2O3 substrates are demonstrated. The β-(AlxGa1−x)2O3 epilayer has an Al composition of 21% and a nominal Si doping of 2 × 1017 cm−3 grown by molecular beam epitaxy. Pt/Ti/Au has been employed as the top Schottky contact, whereas Ti/Au has been utilized as the bottom Ohmic contact. The fabricated devices show excellent rectification with a high on/off ratio of ∼109, a turn-on voltage of 1.5 V, and an on-resistance of 3.4 mΩ cm2. Temperature-dependent forward current-voltage characteristics show effective Schottky barrier height varied from 0.91 to 1.18 eV while the ideality factor from 1.8 to 1.1 with increasing temperatures, which is ascribed to the inhomogeneity of the metal/semiconductor interface. The Schottky barrier height was considered a Gaussian distribution of potential, where the extracted mean barrier height and a standard deviation at zero bias were 1.81 and 0.18 eV, respectively. A comprehensive analysis of the device leakage was performed to identify possible leakage mechanisms by studying temperature-dependent reverse current-voltage characteristics. At reverse bias, due to the large Schottky barrier height, the contributions from thermionic emission and thermionic field emission are negligible. By fitting reverse leakage currents at different temperatures, it was identified that Poole–Frenkel emission and trap-assisted tunneling are the main leakage mechanisms at high- and low-temperature regimes, respectively. Electrons can tunnel through the Schottky barrier assisted by traps at low temperatures, while they can escape these traps at high temperatures and be transported under high electric fields. This work can serve as an important reference for the future development of ultrawide bandgap β-(AlxGa1−x)2O3 power electronics, RF electronics, and ultraviolet photonics.
I. INTRODUCTION
Ultrawide bandgap (UWBG) semiconductors have garnered considerable attention in recent years due to their promising applications in power electronics, optoelectronics, and RF electronics.1–8 β-Ga2O3 is a promising candidate for UWBG semiconductors due to a high breakdown field of 8 MV/cm and a large bandgap of 4.6–4.9 eV, and a high Baliga's figure of merit (BFOM) compared with GaN and SiC.9 Furthermore, due to the availability of large native substrates, β-Ga2O3 has great potential for cost-effective high-voltage power electronics.10 Moreover, alloying Ga2O3 with Al2O3 can produce (AlxGa1−x)2O3 with increased bandgap (e.g., 4.8–6.2 eV for x = 0 to 0.71).11 β-(AlxGa1–x)2O3 is a monoclinic ternary alloy that is expected to have a higher BFOM than Ga2O3, making it more suitable for power electronic applications.3 Recently, there have been several optoelectronic and power devices demonstrated using β-(AlxGa1–x)2O3. Chen et al.12 reported the effect of O2 concentration on sputtered (AlxGa1–x)2O3 films for deep UV photodetectors. β-(AlxGa1−x)2O3/Ga2O3 heterojunction modulation-doped field effect transistors (MODFETs) have been extensively studied.13–19 Zhang et al.15 reported record low-temperature mobility of ∼2700 cm2/V s in the β-(AlxGa1−x)2O3/Ga2O3 heterostructure. Furthermore, Okumura et al.13 demonstrated β-(AlxGa1−x)2O3 metal–semiconductor field effect transistors (MESFETs). However, most of these power devices are lateral devices where the currents flow laterally, and voltages are handled laterally. These lateral devices usually underperform compared with their material limits. In commercial Si and SiC power devices, vertical architecture dominates, especially for high-voltage high-power applications, due to larger current and voltage handling capability, avalanche capability, no surface-related issues, better heat dissipation, and smaller chip area.
To date, there are very few reports of vertical β-(AlxGa1−x)2O3 devices. One of the major challenges is growing high-quality β-(AlxGa1−x)2O3 layers with high Al contents [e.g., >20% on (010) β-Ga2O3 substrates].13 This is because β-Ga2O3 crystallizes in a monoclinic structure, and Al2O3 prefers a conundrum structure, leading to phase separation at high Al contents.20 Furthermore, due to the lattice mismatch with sufficiently high Al composition, epitaxial growth of β-(AlxGa1−x)2O3 on β-Ga2O3 is challenging. Moreover, Schottky and Ohmic contact behavior on β-(AlxGa1−x)2O3 needs to be further understood comprehensively. However, Ohmic and Schottky contact behavior on β-Ga2O3 is extensively investigated in the past few years.21,22 Jadhav et al.23 investigated temperature-dependent barrier height inhomogeneity in β-Ga2O3 Schottky barrier diodes (SBDs). Ahmadi et al.20 found Schottky barrier heights of Ni to β-(AlxGa1−x)2O3 with different Al compositions, which was attributed to the lateral fluctuation of Al alloy composition. Furthermore, β-(AlxGa1−x)2O3 power devices suffer from low breakdown voltages and high leakage current. A systematic investigation of the leakage mechanism and the temperature-dependent characteristics of β-(AlxGa1−x)2O3 devices are still lacking. In this work, we demonstrate vertical β-(AlxGa1−x)2O3 (x = 0.21) SBDs on free-standing (edge-defined film-fed grown) highly doped (010) β-Ga2O3 substrates and systematically investigate their temperature-dependent forward and reverse electrical characteristics. Different models have been used to comprehensively understand the leakage mechanisms of the vertical β-(AlxGa1−x)2O3 SBDs. This work can serve as an important reference for the design of high-voltage high-power vertical β-Alx(Ga1−x)2O3 power devices.
II. EXPERIMENT
200 nm β-(AlxGa1−x)2O3 (x = 0.21) layer was grown using molecular beam epitaxy (MBE) on the edge-defined film-fed grown (010) β-Ga2O3 substrate (Fig. 1) by Novel Crystal Technology, Inc, Japan. During the MBE growth, Ga and Al are evaporated from effusion cells in RF-generated oxygen plasma to form β-(AlxGa1−x)2O3 thin films.24,25 The Al composition of the grown film can be estimated by the shift of β-Ga2O3 and β-(AlxGa1−x)2O3 (020) peak from XRD spectra.26 More details about the MBE growth and Al content determination can be found elsewhere.24–26 The β-Ga2O3 substrate is heavily doped with [Sn] = 3.1 × 1018 cm−3, while Si doping concentration in β-(AlxGa1−x)2O3 layer was estimated to be 2.0 × 1017 cm−3. Atomic force microscopy (AFM) suggested a film surface RMS roughness of 2.6 nm [Fig. 1(d)], and x-ray diffraction (XRD) confirmed the (020) β-(AlxGa1−x)2O3 peak [Fig. 1(b)] of the film. The device fabrication started with sample cleaning using acetone, isopropyl alcohol (IPA), and distilled water. Then, a 200 nm Ni hard mask was deposited by electron beam (E-beam) evaporation on the film using standard photolithography and liftoff. This Ni hard mask protects the etching of β-(AlxGa1−x)2O3 during mesa-isolation of the devices. Next, SF6 inductive-coupled plasma reactive ion etching (ICP-RIE) was performed at 400 W until a mesa depth of 300 nm was obtained. The Ni hard mask was removed using a Ni etchant, followed by sample cleaning using HF and H2SO4 to remove the etching damage induced by the ICP-RIE dry etching. Then, Ti/Au (20/130) nm cathode (Ohmic contact) was deposited at the back side of the β-Ga2O3 substrate using E-beam evaporation followed by 500 °C rapid thermal annealing (RTA) in N2 environment. Finally, Pt/Ti/Au (20/10/120) nm anode (Schottky contact) was deposited by E-beam evaporation and formed using standard photolithography and liftoff. Figure 1(a) shows the schematic of the fabricated vertical β-(AlxGa1−x)2O3 SBDs on the β-Ga2O3 substrate. The diameter of the SBD (ϕ) is 100 μm. Current density-voltage (J–V) measurements of the fabricated devices were measured using a probe station with a hotplate and a Keithley 2470 source meter. Capacitance-voltage (C–V) measurements were performed on a Hewlett-Packard 4275A LCR meter.
III. RESULTS AND DISCUSSION
Figures 2(a) and 2(b) show the forward characteristics of the vertical β-(AlxGa1−x)2O3 SBD at room temperature. The device exhibited a turn-on voltage of ∼1.5 V, a specific on-resistance of 3.4 mΩ cm2 at +5 V, and an on/off ratio of ∼109. The C–V measurements were performed at room temperature using 1 MHz frequency. Figure 2(c) indicates a carrier concentration of 2.8 × 1017 cm−3 in the β-(AlxGa1−x)2O3 film, which is close to the nominal Si doping. The temperature-dependent forward J–V characteristics of the device are presented in Fig. 2(d). The measured electrical characteristics are quite stable and reproducible among many devices across the wafer. The devices after high-temperature testing retain the initial J–V curves even after cooling down to room temperature. Using the diode thermionic emission (TE) model, the Schottky barrier height and ideality factor of the vertical β-Alx(Ga1−x)2O3 SBD can be calculated. For , the TE model can be expressed as27
where J is the current density, is the saturation current density, is the Richardson constant, T is the temperature in kelvin, q is the electron charge, is the effective Schottky barrier height, n is the ideality factor, k is the Boltzmann constant, is the effective electron mass, and h is the Planck constant.
Figure 3 shows the Schottky barrier heights and ideality factors of the vertical β-(AlxGa1−x)2O3 SBD calculated using the TE model. It should be noted that the Richardson constant of 37.8 A/cm2 K2 was calculated using the reported effective electron mass of 0.313m0 (Ref. 3) for β-(AlxGa1−x)2O3. With increasing temperature, varied from 0.91 to 1.18 eV, while n changed from 1.8 to 1.1 [Fig. 3(a)]. There is a clear temperature dependence of both parameters, which stems from the inhomogeneous metal/semiconductor interface.28,29 In Fig. 3(b), a linear correlation between and n was observed, which is a well-known phenomenon in the presence of an inhomogeneous Schottky contact. In the metal/semiconductor interface, there are regions with low and high Schottky barrier heights. At low temperatures, electrons can only pass through low Schottky barrier height regions, whereas at high temperatures, electrons gain momentum to cross high Schottky barrier height regions. As a result, the Schottky barrier height increased with temperature. To further investigate this behavior, the Schottky barrier height can be considered a Gaussian distribution of potential with a mean barrier height and a standard deviation , and the barrier is linearly dependent on voltage as follows:
where and are the values at zero bias, and the coefficients and represent the voltage-induced deformation of the Schottky barrier distribution.30 Substituting Eqs. (5) and (6) into Eq. (4) and combining Eqs. (1) and (2), the ideality factor can be written as31–33
Therefore, Eq. (7) implies ideality factor becomes temperature-dependent, and its value can exceed unity. The voltage dependence of the Schottky barrier height and ideality factor was attributed to the interfacial states at the metal/semiconductor interface. These states become more negative with applied forward bias, leading to an increase in the Schottky barrier height with bias and ideality factor greater than unity.32 This voltage dependency of the Schottky barrier height can be interpreted as image force shifting the Schottky barrier maxima away from the metal–semiconductor into the semiconductor as the forward bias increases.33 Furthermore, theoretical derivations suggest that both and are linear functions of inverse temperature , which is in good agreement with the experimental results in Figs. 3(c) and 3(d). The extracted and were 1.81 and 0.18 eV, respectively. It is reported that Pt on (010) β-Ga2O3 has an ideal Schottky barrier height of up to 1.93 eV.34 This work shows the mean Schottky barrier height at zero bias is 1.81 ± 0.18 eV, which is consistent with the ideal value.
It should be noted that the C–V measurements did not correctly estimate the barrier height from Fig. 2(c). This discrepancy is likely due to two reasons. First, there is frequency dispersion in the C–V measurements in WBG semiconductors. Due to this, the frequency that uses for the C–V measurements may not be the ideal frequency to measure the barrier height. At a particular frequency, some impurity states or donor states respond slowly to the applied AC electric field. To study this effect, frequency dispersion C–V needs to be measured, which is out of the scope of this study. Second, the interfacial states between the metal and the semiconductor are insensitive to C–V measurements at low voltages. The inhomogeneity at the metal/semiconductor interface lowers the accuracy of the C–V measurements, which increases the uncertainty of the measured barrier height. However, barrier height measured from the J–V measurements are quite reliable as it represents the current flowing through the inhomogeneous Schottky barrier.
Figure 4(a) shows the temperature-dependent reverse J–V characteristics of the vertical β-(AlxGa1−x)2O3 SBD. The reverse leakage current increased with increasing temperature. Figure 4(b) represents the current density as a function of 1/T at different reverse voltages. Different gradients in Fig. 4(b) corresponded to different leakage mechanisms. Figure 4(c) shows potential leakage mechanisms for SBDs. Different temperature regimes activate different processes inside the semiconductor, which can be investigated from the reverse J–V measurements. Due to the high Schottky barrier height of >1 eV in the vertical β-(AlxGa1−x)2O3 SBD, TE or thermionic field emission (TFE) is unlikely to be a major contributor to the reverse leakage of the device. The possible candidates for the vertical β-(AlxGa1−x)2O3 SBD include Poole–Frenkel emission (PFE), trap-assisted tunneling (TAT), Fowler–Nordheim tunneling (FNT), field emission (FE), and variable range hopping (VRH). More details about these processes can be found elsewhere.35–41 To identify the dominant process, fitting the reverse leakage data using the mathematical expression of each model is widely used. It was found that two dominant mechanisms were PFE and TAT, while the other mechanisms played a minor role.
The PFE is a trap-mediated transport mechanism where the carrier density depends exponentially on the activation energy of the traps, and the current density due to PFE is given by41
where E is the electric field in the β-(AlxGa1−x)2O3 layer, is the barrier height for electron emission from a trap state, is the high-frequency relative dielectric permittivity, is the permittivity of free space, C is a proportionality constant, and k is the Boltzmann constant. It should be noted that the high-frequency (optical) dielectric constant, rather than the static one, should be used in the PFE equation.41 Electrons can move slowly through an insulator or a semiconductor in the presence of a large electric field. Initially, electrons are in a localized or trap state and cannot move freely. With a high electric field, these localized electrons can be promoted to the conduction band and contribute to leakage currents. These electrons can move through the crystal before relaxing into another localized state. In other words, the PFE model describes an electric-field enhanced emission from a trap state which increases reverse leakage. From Eq. (8), is a linear function of ,
Figure 5(a) indicates the PFE plot for the vertical β-(AlxGa1−x)2O3 SBD between 373 and 473 K. The transport model based on the PFE in Eq. (9) shows a good agreement with the experimental data in this temperature regime with consistent Schottky barrier height and dielectric constant. Figure 5(b) shows the temperature-dependent slope and intercept extracted from Fig. 5(a). Both and are linear functions against inverse temperature, which is expected in the PFE model. The extracted dielectric constant and the emission barrier height for the device were and , which are consistent with experimental results.1,14,42 The obtained is comparable with the extracted values from the TE model. These results indicate that the PFE mechanism is dominant at high temperatures (>373 K).
Below 373 K, the leakage current of the vertical β-(AlxGa1−x)2O3 SBD deviated from the PFE model and leaned toward the TAT model. In the TAT model, an electron in the metal could be activated to a trap state at the metal/semiconductor interface and then tunnel to the semiconductor side.43,44 This model can be expressed by
where is the barrier height for electron emission from a trap state, is effective electron mass, C is a proportionality constant, and is the reduced Planck's constant. The equation can be rearranged to find the from the measured reverse leakage current,
Figure 5(c) shows the measured and theoretical TAT plots for the vertical β-(AlxGa1−x)2O3 SBD between 298 and 353 K. The extracted barrier height for electron emission from a trap state was between 0.92 and 1.1 eV. At low temperatures, electrons can tunnel through the Schottky barrier height with the assistance of traps. With enough thermal energy at high temperatures, they can escape from the trap levels and transport under high electric fields (i.e., the PFE model). For high-voltage applications, growing thick β-(AlxGa1−x)2O3 is needed and demands further research and development.
IV. CONCLUSIONS
We demonstrated vertical β-(AlxGa1−x)2O3 SBDs on free-standing (010) β-Ga2O3 substrates. The device exhibited excellent forward rectifying behaviors with a high on/off ratio of ∼109, a turn-on voltage of 1.5 V, and an on-resistance of 3.4 mΩ cm2. Temperature dependence of the Schottky barrier height and ideality factor was due to the inhomogeneous metal/semiconductor interface. This interface inhomogeneity was explained by considering the Schottky barrier height as a Gaussian distribution with a mean Schottky barrier height of 1.81 eV at zero bias with a standard deviation of 0.18 eV. Comprehensive reverse leakage analysis indicated that PFE and TAT mechanisms were the main contributors to the reverse leakage currents of the device. The extracted physical parameters from the PFE and TAT models are consistent with experimental results. This work can serve as an important reference for the future development of β-(AlxGa1−x)2O3 based electronics and photonics for high-power, high-voltage, and ultraviolet (UV) photonic applications.
ACKNOWLEDGMENTS
Research supported as part of the ULTRA, an Energy Frontier Research Center funded by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences (BES), under Award No. DE-SC0021230 (computational study) and by the National Science Foundation (NSF) under Award No. 2302696 (experimental study and data analysis).
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Dinusha Herath Mudiyanselage: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Investigation (lead); Methodology (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (lead). Dawei Wang: Conceptualization (supporting); Data curation (supporting); Formal analysis (supporting); Investigation (supporting); Validation (equal). Houqiang Fu: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Funding acquisition (lead); Investigation (lead); Methodology (lead); Project administration (lead); Supervision (lead); Writing – original draft (lead); Writing – review & editing (lead).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.