Introduction

Power devices are essential building blocks for high-efficiency energy conversion in power electronics systems. The market size of power semiconductor devices has reached US$40 billion driven by applications like electric vehicles, data centers, electric grids, and renewable energy processing1. Deployment of new semiconductors is a fundamental driving force to advance power electronics. The last decade witnesses the success of wide bandgap (WBG) semiconductors, e.g., gallium nitride (GaN) and silicon carbide (SiC)2,3,4,5. On the horizon, ultrawide bandgap (UWBG) semiconductors hold tremendous promises for the next-generation power electronics6,7,8,9.

Power devices operate as switches between the high blocking voltage and high conduction current. Their robustness against overvoltage and overcurrent stresses is as important as their performance under normal operations. Such robustness is crucial for any power device, as they allow devices to temporarily survive the common faults in power systems, e.g., short circuit, excessive load, arc/ground faults, before the protection circuitry intervenes10. Avalanche is the desirable mechanism of power devices to withstand overvoltage stresses, as it allows them to accommodate high avalanche current (IAVA) at the avalanche breakdown voltage (BVAVA) and thus dissipate the excessive energy in circuits11. The avalanche and surge current capabilities usually represent the power device robustness against electrical and electrothermal overstress12.

The native p-n junction is the enabling device structure for avalanche and surge robustness in silicon (Si), silicon carbide (SiC), and gallium nitride (GaN) devices12,13, the three power semiconductor technologies have reached commercialization. The avalanche hinges on the impact ionization (I. I.) and multiplication occurring at the junction, as well as the efficient removal of the I. I.-produced non-equilibrium carriers. Under forward bias, the capability to withstand high surge current relies on the high-level carrier injection across the p-n junction, which reduces the device resistance as well as suppresses the ramp-up of power loss and junction temperature12. In contrast to homojunctions, avalanche or surge current robustness has not been demonstrated in heterojunction-based power devices to date. Only a few avalanche-capable heterojunctions were reported in low-voltage, low-power optoelectronic devices, and the avalanche-like features observed in some unipolar heterostructures are still in debate14.

Compared to Si and WBG counterparts, UWBG power devices possess superior performance limits1,8. However, native p-n homojunction is difficult to form in UWBG materials due to the challenges of achieving efficient bipolar doping15. This fundamentally limits the robustness of UWBG power devices, of which Gallium oxide (Ga2O3) is an example16,17,18. Benefitting from its high critical electric field, controllable n-type doping, and the large-area wafer availability, Ga2O3 power devices are advancing fast towards applications19,20, Whereas, due to the flat valence band and strong self-trapping of holes, the reliable p-type doping in Ga2O3 is very challenging, although p-type Ga2O3 has been reported by some group21,22,23. As an alternative, heterojunctions between Ga2O3 and foreign p-type oxide, e.g., nickel oxide (NiO) or copper oxide24, have recently been deployed in the design of Ga2O3 bipolar power devices15. Despite excellent device performance, the viability of avalanche and surge robustness in such heterojunctions remains a fundamental knowledge gap. Meanwhile, the impact of band discontinuity on carrier transport is largely unexplored under the high electric field (E-field), high current density, and fast switching conditions.

This work fills this gap by demonstrating avalanche and surge robustness in NiO/Ga2O3 p-n heterojunctions through device innovations and circuit characterizations, whilst relevant carrier dynamics are also revealed through microscopic techniques and physics-based simulations. Large-area NiO/Ga2O3 p-n heterojunction diodes (HJDs) with advanced edge terminations are designed and fabricated, followed by the avalanche and surge circuit tests complying with industrial standards. Subsequently, the electron beam-induced current (EBIC) characterization and simulations reveal the carrier transport dynamics under critical avalanche and surge conditions. As a key enabler for surge robustness, the bipolar conductivity modulation is found to be dominantly in the NiO with a high hole concentration, while it is usually in the lightly-doped side in conventional homojunctions. This distinction allows NiO/Ga2O3 heterojunctions to simultaneously achieve a smaller reverse recovery and higher switching speed with robustness comparable or superior to that of conventional homojunction.

Results

NiO/Ga2O3 p-n heterojunction diode

For power devices, edge termination design is critical to control the E-field crowding, avoid premature breakdown, and access the device BVAVA. Here we employ an edge termination that combines small-angle beveled junction termination extension (JTE) and a high-k field plate. Figure 1a presents the three-dimensional schematic diagram of the large-area (3 mm × 3 mm) NiO/Ga2O3 p-n HJD fabricated on 2-inch free-standing Ga2O3 wafers. The p-type region consists of lightly- and heavily-doped NiO layers (i.e., p-NiO and p+-NiO). The p-NiO layer can reduce the leakage current and favor the JTE design. At the device edge, the p-NiO extension functions as a JTE, and the small beveled angle allows for a gradual decrease in charge density away from the active region, which continuously reduces the depletion curvature and surface E-field25. The high-k field plate conformally covers the NiO JTE and can further passivate the peak E-field.

Fig. 1: NiO/Ga2O3 p-n heterojunction power device.
figure 1

a Three-dimensional schematic of the NiO/Ga2O3 HJD, showing the double-layered NiO (a 300-nm-thick p-NiO and a 100-nm-thick p+-NiO) and a high permittivity BaTiO3 dielectric layer with 11° beveled-mesa termination. b STEM image of the termination region of the heterojunction in false color to highlight different layers. c High-resolution cross-section TEM image of the heterojunction interface. d Schematic energy band diagram of the HJD at zero bias. e Simulated in-plane E-field contour of devices with and without the BaTiO3 dielectric layer. f Frequency-dependent capacitance characteristics measured at different reverse bias voltages.

The HJD fabrication starts with the deposition of NiO films on n-Ga2O3 drift layer via CMOS-compatible RF magnetron sputtering technique. The hole concentrations in NiO are modulated by tuning the gas flux ratios of Ar/O2 in the sputtering process, resulting in a hole concentration of 5.8 × 1017 cm−3 and 2.9 × 1019 cm−3 in p-NiO and p+-NiO layers, respectively. The beveled angle in NiO is implemented by adjusting the gap between the shadow mask and Ga2O3 wafer as well as the declination angle of the NiO target in the sputtering process, as detailed in Supplementary Section S1. Amorphous barium titanate (BaTiO3), a perovskite oxide with an ultrahigh dielectric constant, is deposited by RF sputtering between the anode metal and NiO to form the field plate. In addition to bare-die devices, some HJDs are sealed in TO-220 packages for circuit tests. More detailed fabrication and packaging process is described in Methods, Supplementary Section S1, and Supplementary Movie 1. Figure 1b shows the cross-sectional scanning transmittance electron microscopy (STEM) image of the edge termination, revealing a bevel angle of 11o. The high-resolution TEM image in Fig. 1c shows an atomically sharp interface of the NiO/Ga2O3 junction with excellent lattice alignments of the (111)-oriented NiO with (001) Ga2O3.

TCAD simulations are performed to investigate the heterojunction band structure and the device E-field management. The simulation models are detailed in Supplementary Section S4. As shown in Fig. 1d, at equilibrium, the NiO/Ga2O3 junction exhibits a type-II (staggered) band alignment with the conduction band and valence band offsets being 2.1 and 3.2 eV, respectively. Figure 1e shows the simulated E-field contours in the HJD with only the NiO JTE and full termination, both at a reverse bias of 1600 V. The high-k field plate shifts the peak E-field away from the junction edge and lowers the peak E-fields in Ga2O3 and NiO (from 6.61 MV/cm to 4.57 MV/cm and from 7.91 MV/cm to 2.62 MV/cm, respectively). As a result, a nearly uniform E-field is present at the NiO/Ga2O3 junction, enabling a uniform and robust avalanche.

Capacitance-voltage (C–V) characterizations are performed for the HJD at frequencies ranging from 1 kHz to 1 MHz, showing negligible frequency dispersion at various biases (Fig. 1f). This indicates the presence of minimal interface states at the heterojunction. The net donor concentration in the Ga2O3 drift layer and the built-in potential of the heterojunction are also extracted from the C–V characteristics to be 1.7 × 1016 cm−3 and 2.1 V, respectively (see Supplementary Section S2). From the built-in potential and band offsets, the barrier heights for electrons and holes are 4.2 and 5.3 eV, respectively. This suggests the electron injection could be more pronounced than hole injection at large forward biases.

Avalanche breakdown robustness

Avalanche breakdown is desirable for both power devices and power electronics systems. For devices, it allows for a non-destructive breakdown with a positive temperature coefficient of BVAVA12 and a smaller overvoltage margin as required for a certain voltage rating26. For systems, the concurrence of high IAVA and high BVAVA can dissipate the surge energy and prevent it from further circulating in the circuitry11. The avalanche characterization should cover all these signatures. Here we employ the quasi-static current-voltage (I–V) characterization to probe the BVAVA behaviors and the unclamped inductive switching (UIS) circuit to measure the IAVA and avalanche energy (EAVA)27.

Figure 2a shows the reverse I–V–T characteristics of the NiO/Ga2O3 HJD at temperatures (T) from 25 to 175 °C. The reverse current is low (<0.2 µA) and shows weak bias dependence below 1500 V. Then it rises sharply due to the initiation of I. I. and multiplication. The BVAVA increases from 1545 to 1683 V with T elevating from 25 to 175 °C, showing a positive temperature coefficient of 1 V/°C. As compared to a similarly-rated commercial SiC diode with a leakage current of ~10 µA under a reverse bias of 1200 V28, the NiO/Ga2O3 HJD shows 102~103 times lower leakage current at high bias, reflecting the superior properties of UWBG materials.

Fig. 2: Robust avalanche breakdown and the relevant carrier dynamics.
figure 2

a Temperature-dependent reverse I–V characteristics of the HJD. b Photograph of the UIS test setup and its circuit diagram. c Typical temperature-dependent UIS voltage and current waveforms for the Ga2O3 HJD at a LUIS of 1 mH. d Temperature coefficient values (1.0 V/°C) for BV extracted from quasi-static and UIS measurements, respectively. e IAVA and avalanche energy (EAVA) as a function of LUIS. f Illustration of the carrier transport dynamics under the avalanche condition. g Simulated contour of the impact ionization (I. I.) generation rate at BVAVA. h, i Simulated profiles of the electron and hole concentration, E-field, and I. I. generation rate in the HJD under two different hole mobilities, at an IAVA of 30 A.

The UIS setup and test procedure follows the industrial JEDEC standard29, with the details elaborated in Supplementary Section S3. Figure 2b shows the UIS circuit schematic and the prototyped test board. In the UIS test, the switch (a power transistor with BV higher than that of the HJD) is first turned ON to charge the inductor (LUIS); the switch is then turned OFF, forcing the HJD to withstand the surge energy stored in LUIS. Figure 2c shows the test waveforms at T from 25 to 175 °C, revealing the desired avalanche waveforms: the HJD voltage clamps at BVAVA with the decreasing of IAVA from 30 A to zero, and the energy stored in LUIS is fully dissipated in the HJD in this 20 μs avalanche process. Repeating this avalanche process, Fig. 2c presents the sustainability of the HJD under 1-million cycles of repetitive avalanche tests. The device forward and reverse I–V characteristics before and after this cycle test show minimal parametric shifts, as presented in Supplementary Section S5. In addition, the BVAVA extracted from the UIS waveforms shows a temperature coefficient identical to that extracted from the I–V characteristics (Fig. 2d). Note the BVAVA in the UIS test (1740 V) is higher than that in the I–V characteristics, due to a higher junction temperature in the UIS test under high IAVA.

The UIS tests are then performed under various LUIS and the relevant charging time, both of which can alter IAVA and EAVA. Avalanche waveforms are obtained under all these conditions. Figure 2a shows the expanded I–V characteristics that combine the quasi-static I–V curves (at low-current levels) and the IAVA ~ BVAVA data obtained from the UIS tests. Both the HJD and the reference SiC diode show a smooth transition between the two sets of data, suggesting a consistent avalanche across the main p-n junction under a wide range of IAVA (inconsistent avalanche locations would lead to abrupt transition27). Figure 2e shows the IAVA and EAVA under three LUIS, revealing an IAVA up to 50 A and EAVA up to 730 mJ.

Critical band structure and carrier dynamics are scrutinized to understand the robust avalanche in the HJD. As shown in Fig. 2f, after I. I. is initiated in n-Ga2O3, the high E-field sweeps the produced electrons and holes to the cathode and the heterojunction, respectively. Under reverse bias, the staggered band produces no barriers for hole transport, which enables efficient hole removal (and high IAVA). From the expanded avalanche I–V characteristics (Fig. 2a), the I. I. coefficients of electrons (αn) and holes (αp) in Ga2O3 are calculated. Note that only a theoretical αn30 and no αp have been reported for Ga2O3 previously. The calculated αn and αp are then fed into the TCAD simulation based on the Selberherr I. I. model31. The calculation and simulation models are detailed in Supplementary Section S4. Figure 2g shows the simulated contour of the I. I. generation rate at BVAVA. The peak I. I. is located in Ga2O3 near the junction and uniformly distributed in the lateral dimension, confirming I. I. initiation in Ga2O3 and avalanche process across the entire device active region.

The demonstration of avalanche in Ga2O3, which has not been reported previously, also sheds light on the minority carrier (hole) transport in Ga2O3, which remains controversial32,33, and largely unexplored under high E-field. As high IAVA hinges on efficient hole removal, the I. I.-produced holes in Ga2O3 are believed to be exempt from self-trapping under high E-field and drift with considerable mobility (μp). To estimate μp, we simulate the avalanche dynamics with μp of 10−4 and 1 cm2 V−1s−1 according to the low and high values theoretically predicted in the literature32,34, (μp is the low-field hole mobility, and field-dependent mobility model is detailed in Supplementary Section S4). Note that higher μp values, e.g., ~1.2 cm2/V s34 and 8~10 cm2/V s22,35, have been reported experimentally. Here we use two lower μp values in the simulation mainly to consider the worst scenario of avalanche, as a high μp can allow for a more efficient hole removal and thus supports a high avalanche current. At BVAVA, the drift velocity does not reach saturation under either μp values. Figure 2h, i shows the simulated contours of carrier concentrations, E-field and I. I. generation rate for the two μp, respectively. The low μp would induce serious hole accumulation and high E-field crowding at the heterojunction, making it unlikely to sustain a stable avalanche. These effects are eliminated for μp = 1 cm2 V−1s−1, rendering it a more reasonable μp to explain the avalanche in n-Ga2O3.

Surge current and reverse recovery characteristics

While avalanche represents the HJD’s robustness at reverse bias, surge current measures its capability to withstand forward overcurrent. Here, a 10-ms half-sinusoidal current pulse with an adjustable amplitude is employed for surge current characterization following the JEDEC standard36. Figure 3a shows the circuit schematic and the prototyped setup, with the circuit design detailed in Methods. Figure 3b, c shows the current and voltage waveforms in the surge current tests with increased amplitude. The HJD can withstand over 50 A surge current, under which condition the forward voltage approaches 14 V. Based on these time-resolved data, the surge current I–V locus are plotted in Fig. 3d for the HJD and a reference Ga2O3 Schottky barrier diode (SBD) fabricated on the same wafer. The surge current and voltage characteristics of the reference Ga2O3 SBD are shown in Supplementary Section S6. The locus of the HJD and SBD shows an anticlockwise and clockwise signature, respectively, which signifies negative and positive temperature coefficients (ηT) of the differential on-resistance (RON)12,37, The RON’s negative ηT in the HJD is further confirmed by the forward I–V characteristics measured in the DC mode at elevated T (Fig. 3e) and in the pulse mode with various pulse widths (Fig. 3f). The differential RON decreases with the increased pulse width and forward voltage.

Fig. 3: Surge current and reverse recovery characteristics.
figure 3

a Photograph of the surge current test setup and its circuit diagram. b, c Surge current and voltage waveforms of the HJD, respectively. d Surge I–V locus of the HJD and the reference Ga2O3 SBD, each set of locus with various amplitudes. e, f Quasi-static temperature-dependent I–V characteristics and pulse I–V characteristics of the HJD, respectively. g Photograph of the reverse recovery test setup and its circuit diagram. h reverse recovery characteristics of the Ga2O3 HJD, the reference Ga2O3 SBD, and a commercial Si fast-recovery diode (FRD). i reverse recovery waveforms of the HJD with different conditions.

While a positive ηT of RON is expected for the unipolar SBD due to the mobility drop at high T, the negative ηT suggests bipolar conductivity modulation in the HJD. Bipolar conduction in homojunction power devices is usually dominated by the high-level carrier injection into the lightly-doped drift region, e.g., hole injection into n--type layer, which typically lead to the considerable reverse recovery38. When bipolar devices fast switches from a high forward current to high reverse bias, a wide depletion region has to be established in the lightly-doped drift region, requiring the removal of minority carriers that previously fill this region. This leads to a usual trade-off between surge robustness and switching speed for bipolar power devices.

To explore if a similar trade-off holds for the NiO/Ga2O3 heterojunction, reverse recovery characterization is performed for the HJD, the reference Ga2O3 SBD, and a similarly-rated commercial Si fast-recovery diode (FRD)39, with the circuit schematic and prototype shown in Fig. 3g and detailed in Methods. As shown in Fig. 3h, i, the 1200-V reverse recovery of the HJD is similar to the unipolar Ga2O3 SBD and much faster than the bipolar Si FRD. Moreover, the reverse recovery waveforms are nearly independent of the forward current (5–10 A) and temperature (25–175 °C), suggesting the reverse recovery waveform is dominated by the capacitive ringing instead of minority carrier recombination. The recombination process, if any, should be faster than the duration of the first ringing. Based on the method in ref. 12 and the reverse recovery time (trr) of 12.79 ns extracted from the first ringing waveform, the maximum hole lifetime in Ga2O3 is estimated to be 6.50 ns.

Microscopic EBIC characterization

The high surge current capability albeit minimal reverse recovery in the HJD suggests a different origin of conductivity modulation as compared to conventional bipolar devices. We suspect the origin to be the electron injection into NiO. Under low forward bias, trap-assisted tunneling dominates the interfacial recombination, while under the elevated forward bias, high-level minority carrier (electrons) injection into the NiO layer occurs via tunneling or thermionic emission across the interfacial barrier at the conduction band. Injected electrons diffuse in the neutral region of NiO and contribute to current conduction, reducing the whole RON of the HJD40. Under reverse bias, as the depletion mainly occurs in the lightly-doped Ga2O3, minimal minority carries need to be recombined in p-NiO for switching to occur. Hence, the carrier dynamics in NiO insignificantly impact the device reverse recovery.

To pursue the microscopic visualization of carrier diffusion, we apply microscopic EBIC characterization, an SEM-based technique widely used to image carrier dynamics to the NiO/Ga2O3 junction41. The EBIC signal is recorded simultaneously with the SEM detector signal, thereby allowing us to spatially correlate the EBIC signal. Figure 4a shows the cross-sectional SEM image of the HJD, and the corresponding EBIC map is provided in Fig. 4b. In the presence of built-in E-field in the depletion region, the EBIC signal decays from the junction interface to both sides of the junction, where a higher current signal is observed in NiO. This indicates more minority carriers are injected into NiO, since only minority carriers contribute to the induced current42. The exponential decay of the EBIC signals is well described in terms of I = I0ex/L, where I0 is the maximum intensity and x is the distance from the junction41. Accordingly, the minority carrier diffusion length (L) for electrons in p-NiO and holes in n-Ga2O3 are determined to be 413 and 127 nm, respectively (Fig. 4c). The difference to the reported values in43,44, could be result of the variation in epitaxial quality, synthesis methods, and material conductivities. Consequently, the minority carrier lifetime (τ) for electrons in p-NiO and holes in Ga2O3 are extracted to be 124.0 and 6.2 ns, respectively, based on the relation L = (D × τ)1/2 and the carrier diffusion coefficient (D) in respective materials45. The hole lifetime in Ga2O3 is consistent with the maximum value estimated from the reverse recovery waveform. In addition, in terms of Einstein’s relation, the hole mobility in Ga2O3 is estimated to be 1 cm2 V−1s−1, which is consistent with the value estimated from the avalanche simulations shown in Fig. 2h, i.

Fig. 4: EBIC characterization and the carrier dynamics under the surge condition.
figure 4

a False color cross-sectional scanning electron microscopy (SEM) image of the HJD. b The corresponding colored EBIC map of the heterojunction region under zero bias and at room temperature. c EBIC profile extracted by the integration of EBIC intensity map in the vertical direction. d Illustration of the carrier transport dynamics under the high forward current. e Simulated distribution of minority carriers on both sides of the HJD at a forward voltage of 6 V.

The EBIC results reveal that the L, τ, and minority carrier quantity in Ga2O3 are much smaller than those in NiO, making the conductivity modulation occurs predominantly in NiO. Accordingly, the schematic of band structure and carrier dynamics in the HJD at high forward bias is illustrated in Fig. 4d. At the heterojunction, electrons and holes primarily tunnel through the barriers produced by band offsets, whilst undergoing a trap-assisted band-to-band recombination46. Feeding the parameters extracted from EBIC into the simulation, the contours of electrons and holes are shown in Fig. 4e at a forward voltage of 6 V. High concentration of electrons approaching 1020 cm−3 are present in the entire NiO region, while the hole concentration in Ga2O3 is much lower and drops rapidly away from the junction. The simulation results validate the more pronounced minority carrier injection in NiO under the surge current condition.

Discussion

The intriguing physics of the NiO/Ga2O3 heterojunction enables the fabricated HJD to deliver a breakthrough trade-off between device performance and robustness. Achieving low RON, high current, and high BV concurrently is a major pursuit for power device technologies1. As shown in Fig. 5a, the RON and BV trade-off of our HJD is the record-breaking among the reported ampere-class Ga2O3 diodes with forward current >1 A37,47,48,49,50,51,52,53,54,55, In addition, the benchmark of specific on-resistance (Ron,sp) versus BV of the ampere-class Ga2O3 HJD against small-area Ga2O3 power diodes is shown in Supplementary Section S9. Figure 5b benchmarks the EAVA density of our HJD with homojunction-based Si, SiC, and GaN devices with a similar BVAVA. The EAVA density of our HJD is significantly higher than that of Si devices and comparable to the highest values reported in SiC and GaN devices12,27,28,56,57,58,59,60,61,62, In addition, as shown in Fig. 5c, our HJD shows an excellent combination of high surge current and surge energy comparable to the highest values reported in Si, SiC, and GaN devices12,39,60,61,63,64,65,66,67, Finally, as shown in Fig. 5d, our HJD breaks the trade-off between the reverse recovery time and surge energy in Si, SiC, and GaN devices12,28,39,61,63,64,65,66,67, enabling a much lower switching speed and loss, whilst demonstrating the state-of-the-art robustness.

Fig. 5: Benchmarking performance and robustness of the Ga2O3 HJD against commercial Si, SiC, and GaN power diodes.
figure 5

a Ron versus BV benchmark for all the ampere-class Ga2O3 power diodes. b EAVA versus BVAVA benchmark, c Surge energy (ESUR) versus peak surge current (ISUR) benchmark, and d ESUR versus reverse recovery time (trr) benchmark, all for Ga2O3, GaN, SiC, and Si power diodes.

In summary, this work demonstrates a NiO/Ga2O3 heterojunction architecture with high avalanche and surge current robustness, minimal reverse recovery, as well as an excellent trade-off between RON and BV. The robustness of the large-area, packaged HJD is characterized by in-circuit tests following industrial standards, revealing the survival of >1700 V overvoltage and >50 A overcurrent. This robustness is attributable to the effective edge termination design, high-quality heterojunction, staggered band alignment, and conductivity modulation. Benefitting from the asymmetric minority carrier lifetimes, conductivity modulation is dominant in p-NiO, which is visualized through microscopic EBIC characterization and enables minimal reverse recovery in switching applications. The avalanche and surge current characteristics also provide physical insights into the fundamentals of hole transport dynamics in Ga2O3, particularly the parameters including impact ionization coefficient, high-field mobility, and a lifetime of minority carriers. This heterojunction is promising for making robust power devices in semiconductors lacking bipolar doping for applications in electric vehicles, aerospace, grid, and renewable energy processing.

Methods

Epitaxial structure

The epitaxial structure was grown by hydride vapor phase epitaxy (HVPE) on a conductive Sn-doped (001) β-Ga2O3 substrate, consisting of a 10-μm Si-doped β-Ga2O3 drift layer with an electron concentration of 1.7 × 1016 cm−3. Based on the Hall measurements of the controlled NiO samples on semi-insulating substrates, hole mobilities corresponding to p-NiO (5.8 × 1017 cm−3) and p+-NiO (2.9 × 1019 cm−3) layers were determined to be 0.87 cm2/V s and 0.34 cm2/V s, respectively. Schematic of the processing steps is shown in Supplementary Section S1. The current-voltage characteristics of the p-NiO layer are shown in Supplementary Section S8.

Device fabrication

The device fabrication started with the substrate thinning from 640 μm to 150 μm by chemical mechanical polishing process, followed by substrate cleaning via ultrasonic treatment in acetone and soaking. Then, the Ga2O3 epi-wafers were annealed at 500 °C for 5 hours under the O2 ambient to partially compensate the donors in the epi-layer. The back-side Ohmic contact (cathode) was formed by the Ti/Au (20/80 nm) deposition through electron beam evaporation (EBE), followed by rapid thermal annealing at 500 °C for 1 min under N2 ambient. This annealing process has been reported to be able to effectively passivate near-surface defects in Ga2O3 epi-layer68. Note that the inter-diffusion of Sn from substrate and Si from epi-wafer is expected to be negligible at this annealing temperature69,70, Subsequently, by using an angled shadow mask, a 400-nm-thick double-layered NiO film with adjustable bevel angle was deposited on the Ga2O3 drift layer by RF magnetron sputtering technique at room temperature. During the sputtering process, the substrate was rotated at 4 rpm to enhance the film uniformity. The target was high-purity (99.99%) NiO ceramics. To alleviate the damage induced by sputtering plasma, the initial RF power was 50 W, and then increased to 150 W when the NiO thickness was above 20 nm. The distance between the target and the wafer was maintained at 13 cm. The growth pressure was 0.6 Pa in an Ar/O2 mixed ambient, and the flux ratios of Ar/O2 were tuned from 20:1 to 2:1 to modulate hole concentration in the double-layered NiO. A 300-nm-thick BaTiO3 was also deposited by RF magnetron sputtering at the same growth pressure of 0.6 Pa in an Ar/O2 mixed ambient with a flux ratio of 10:1, followed by annealing at 300 °C in oxygen ambient for 1 hour. The contact window was opened by a lift-off process, and Ni/Au (300/200 nm) metal stack was deposited by EBE to form the anode contact, producing an active area of 3 × 3 mm2.

Device package

The device was sealed in a TO-220 package for circuit testing. For the comparison of surge current characteristics, 9-mm2 Ni/β-Ga2O3 SBDs were also fabricated on the same wafer by identical processes except for the p-NiO deposition underneath the Schottky anode.

Device static electrical characterizations

The quasi-static forward/reverse I–V characterizations were performed by a B1505 power device analyzer. The pulse measurement mode in the B1505 analyzer was adopted to characterize the forward current above 1 A (Fig. 3e, f), while the high-resolution DC mode was used to measure the log-scale forward I–V characteristics of the HJD (Supplementary Section S7). The C–V characteristics were measured by using an E4980A precision LCR meter at room temperature.

Device circuit-level characterizations

The UIS, surge current, and reverse recovery characterizations were carried out by a customized circuit test platform. All test circuits and methods were formulated according to the Joint Electron Device Engineering Council (JEDEC) standards. Photographs of test circuits and experimental platforms are presented in Supplementary Section S3.

EBIC and TEM characterizations

EBIC measurement was carried out in an FEI Helios 600 NanoLab Dual-beam FIB system equipped with Kleindiek Nano Control NC40 nanomanipulators and low-current measurement units. The top- and bottom electrodes were contacted to the nanomanipulator and sample stage, respectively, allowing electrical current to flow and pass through a current amplifier. The electron beam direction is perpendicular to the surface of the junction, and the acceleration voltage and electron beam current were 15 kV and 0.17 nA, respectively. Line-scan measurement from the top electrode down to the Ga2O3 drift layer was performed to extract the profile of the EBIC current across the junction. Scanning TEM, bright-field high-resolution TEM, and energy dispersive X-ray (EDX) spectroscopy elemental mapping were performed using an FEI Tecnai F-20 microscope (FEI TF-20), operated at an acceleration voltage of 200 kV.

Simulations

The Technology Computer-Aided Design (TCAD) device simulations were performed using the Silvaco TCAD device package. Additional details for the simulations are provided in Supplementary Section S4.