Next Article in Journal
Effect of Electrochemical Hydrogen Charging on Blistering and Mechanical Properties Behavior of Q690 Steel
Previous Article in Journal
Synthesis of Thiophene-Based Derivatives and the Effects of Their Molecular Structure on the Mesomorphic Behavior and Temperature Range of Liquid-Crystalline Blue Phases
Previous Article in Special Issue
One-Step Preparation of Si-Doped Ultra-Long β-Ga2O3 Nanowires by Low-Pressure Chemical Vapor Deposition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Breakdown Characteristics of Ga2O3-on-SiC Metal-Oxide-Semiconductor Field-Effect Transistors

1
College of Integrated Circuit Science and Engineering, Nanjing University of Posts and Telecommunications, Nanjing 210023, China
2
National and Local Joint Engineering Laboratory for RF Integration and Micro-Assembly Technologies, Innovation Center for Gallium Oxide Semiconductor (IC-GAO), Nanjing 210023, China
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(6), 917; https://doi.org/10.3390/cryst13060917
Submission received: 10 May 2023 / Revised: 29 May 2023 / Accepted: 5 June 2023 / Published: 6 June 2023
(This article belongs to the Special Issue Wide Bandgap Semiconductor Electronics and Optoelectronics)

Abstract

:
Ultra-wide bandgap semiconductor gallium oxide (Ga2O3) features a breakdown strength of 8 MV/cm and bulk mobility of up to 300 cm2V−1s−1, which is considered a promising candidate for next-generation power devices. However, its low thermal conductivity is reckoned to be a severe issue in the thermal management of high-power devices. The epitaxial integration of gallium oxide thin films on silicon carbide (SiC) substrates is a possible solution for tackling the cooling problems, yet premature breakdown at the Ga2O3/SiC interface would be introduced due to the relatively low breakdown strength of SiC (3.2 MV/cm). In this paper, the on-state properties as well as the breakdown characteristics of the Ga2O3-on-SiC metal-oxide-semiconductor field-effect transistor (MOSFET) were investigated by using the technology computer-aided design (TCAD) approach. Compared with the full-Ga2O3 MOSFET, the lattice temperature of the Ga2O3-on-SiC MOSFET was decreased by nearly 100 °C thanks to the high thermal conductivity of SiC. However, a breakdown voltage degradation of >40% was found in an unoptimized Ga2O3-on-SiC MOSFET. Furthermore, by optimizing the device structure, the breakdown voltage degradation of the Ga2O3-on-SiC MOSFET is significantly relieved. As a result, this work demonstrates the existence of premature breakdown in the Ga2O3-on-SiC MOSFET and provides feasible approaches to further enhance the performance of hetero-integrated Ga2O3 power devices.

1. Introduction

Due to the wide bandgap (~4.8 eV) [1], high Baliga’s figure-of-merit (BFOM) [2], and outstanding breakdown strength (~8 MV/cm) [3], gallium oxide (Ga2O3) has shown great potential in power devices including metal-oxide-semiconductor field effect transistors (MOSFETs) and Schottky barrier diodes (SBDs) [4,5,6]. For high-power devices, thermal cooling efficiency is vital for operation robustness and lifetime, which requires dedicated thermal management [7,8,9]. A major disadvantage of gallium oxide is its low thermal conductivity [10], which, if not well considered, will reduce the performance of Ga2O3-based devices. Device robustness is critical for electronic systems, especially for high-power systems, and thermal management is the key to improving device robustness. System failure may occur due to the harsh radiation, which is more likely to occur when the device is at a high working temperature [11,12,13,14,15,16]. One solution to this problem is packaging Ga2O3-based devices with materials or structures with high thermal conductivity [8]. Another method is fabricating devices on a high-thermal conductivity substrate. Al2O3 (thermal conductivity = 0.3 W∙cm−1K−1) is a suitable substrate material for Ga2O3-based power devices and shares a better thermal conductivity than Ga2O3 (thermal conductivity = 0.13 W∙cm−1K−1). However, compared with Al2O3, silicon carbide (SiC) is an ideal substrate that features a superior thermal conductivity of 4.9 W∙cm−1K−1 and is a material with which Ga2O3 hetero-epitaxy can be realized [17].
It has been revealed that hetero-epitaxial Ga2O3-on-SiC MOSFETs have more advantages in thermal performance than homo-epitaxy [18,19]. However, SiC has a relatively low critical electric field compared with Ga2O3, which could impair the breakdown voltage (BV) due to premature breakdown [20]. Meanwhile, crystal dislocations could be introduced at the interface of two materials, which results in harmful interface states [21]. However, thanks to the progress in epitaxial technology, such a non-ideal factor would be relieved [22,23]. Previous studies mainly focused on the on-state performance of the Ga2O3-on-SiC MOSFET by considering the lattice self-heating effects but with little concentration on breakdown characteristics [18,19,24,25].
The low-cost method of numerical computation has been widely used for the research of material properties, device design, circuit, and system analysis [26,27,28,29,30,31,32]. For SiC-based devices, the convergence of numerical computation has been solved by advances in computing technology [33]. However, for Ga2O3 devices, the poor convergence of numerical computation is currently an issue due to the ultra-wide bandgap of Ga2O3 [34] and the low intrinsic carrier concentration.
In this study, the Ga2O3-on-SiC MOSFET was designed, and its on-state as well as breakdown characteristics were simulated. Performance, including the transfer, output, and electric field distribution, is presented using Synopses Sentaurus TCAD. On-state, lattice self-heating effects were taken into consideration, lattice temperatures were calculated, and their effects on carriers’ mobility were discussed. By comparing the differences between MOSFETs on SiC substrates and Ga2O3 substrates in temperature distribution and on-state performance, the thermal effects of the SiC substrate are verified. Off-state, electric field distribution was adopted to indicate the breakdown voltage in this study [26]. The effect of SiC concentration, SiC thickness, Ga2O3 epitaxial doping concentration, and Ga2O3 epitaxial thickness on the breakdown voltage was evaluated. SiC premature breakdown was discussed, and basic improvements to minimize this nonideal effect are also provided.

2. Device Structure and Simulation Setup

The structures of MOSFETs in this study are shown in Figure 1. In Figure 1a, the conventional Ga2O3 MOSFET consists of a semi-insulating Ga2O3 substrate (thickness = 200 μm) and a Ga2O3 epitaxy layer. In Figure 1b, the Ga2O3 substrate was replaced by a p-type/semi-insulating SiC layer. Below the p-type/semi-insulating SiC layer, the conductive n-type substrate (thickness = 200 μm) and substrate electrode were set for thermal simulation, which are not shown in the figure. For both device structures, thermal contacts were deployed, and the contact temperature was set to 300 K. The source and drain electrodes were treated as ohmic contacts, and the work function of the gate electrode was set at 5.3 eV. The distance between the drain and gate, also known as the drift length in power devices, was set to 7.5 μm, and the distance between the source and gate was 0.5 μm. A thickness of 20 nm of Al2O3 was used as the gate dielectric, and a passivation layer of Si3N4 was also introduced. All parameters are selected based on experimental results, and the simulation has been calibrated [35]. Detailed structural parameters can be found in Table 1.
In this study, carrier mobility was obtained by solving several physical equations in conjunction with each other. The carrier’s mobility is mainly governed by three effects, namely: doping concentration dependence, high-field saturation, and surface effect [36]. In this study, the carriers, i.e., electrons, were distributed in the channel region of the MOSFET. Thereby, the surface effect can be considered to have a relatively minor impact on the carrier’s mobility and thus was ignored. The high electric field saturation effect and the carrier doping concentration effect were considered. Many theoretical approaches were used to calculate mobility. The experimental results were summarized by the Arora model to represent the effect of carrier doping effects on carrier mobility. The corresponding model representation is as follows [37]:
μ doping = μ min + μ d 1 + ( N A , 0 + N D , 0 N 0 )
μ min = A min ( T T 0 ) α m
μ d = A d ( T T 0 ) α d
Ν 0 = A N ( T T 0 ) α N
A * = A a ( T T 0 ) α a
where Amin, Ad, AN, and Aa are the model parameters depending on materials, T is the ambient temperature, and T0 is the room temperature, which was taken as 300 K in this study. ND,0 and NA,0 in Equation (1) are the donor and acceptor concentrations, respectively. μmin is the minimum mobility for materials with a high doping concentration. μd is the differential value between minimum mobility and mobility under low doping conditions. Equation (1) interprets the relationship between doping concentration and carrier mobility. In Equations (1)–(5), when the temperature is fixed, N0, μd, and μmin are all fixed values. With a higher ND,0 or NA,0, the μdoping decreases. This model successfully shows the trend of doping concentration as a function of carrier mobility. The parameters of the Ga2O3 Arora model in this work were extracted from experimental results in references [36,38]. The calibration results were shown in Figure 2, and the extracted parameters were listed in Table 2.
As for the carriers’ high electric field saturation effects, the Canali model was used in this study, which was derived from the Caughey-Thomas model. The mobility can be expressed by Equations (6) and (7), where μlow is the carrier mobility at a low electric field, μlow is determined by the mobility previously obtained at a low electric field condition, and vsat is the carrier saturation velocity. In semiconductors subject to scattering, there is an upper limit to the carrier velocity, and it varies for different materials. The coefficient β is a temperature-dependent quantity that satisfies the following relationship. The Canali model parameters used in this study for Ga2O3 are shown in Table 3 [36,39,40].
μ ( E ) = μ min + ( α + 1 ) μ low α + [ 1 + ( ( α + 1 ) μ low E v sat ) β ] 1 / β
β = β 0 ( T T 0 ) β exp
Meanwhile, in this study, carrier mobility was found to be closely related to lattice temperature. It has been shown that the mobility of the carrier has a negative exponential relationship with the lattice temperature, as expressed in Equation (8), where μ0 represents the electron mobility at 300 K, T represents the lattice temperature, T0 is 300 K, α1 represents the corresponding exponential coefficient, and a value of 2 is typically used for α1 in Ga2O3 [41].
μ = μ 0 ( T T 0 ) α 1
For breakdown characteristics, avalanche breakdown is the key point. In this process, electron-hole pairs were generated by impact ionization. However, due to the ultra-wide bandgap and low carrier concentration of Ga2O3 (10−22 cm−3), it was difficult to directly calculate the reverse current by numerical calculation and often encountered non-convergence. To address this issue, the internal electric field distribution was calculated at different voltages. According to the related literature, the critical breakdown field strengths of Ga2O3 and SiC are 8 and 3.2 MV/cm [1,33], respectively. In this study, the drain voltage was continuously increased to analyze the internal electric field distribution, and the device was considered to break down when either Ga2O3 or SiC reached its breakdown field strength, which is a common method adopted for determining the breakdown voltage of Ga2O3-based devices [26].

3. On-State Characteristics

The transfer characteristics of the Ga2O3-on-SiC MOSFET and the conventional Ga2O3 MOSFET were first investigated. As shown in Figure 3, the transfer characteristics are displayed, where Figure 3a presents the transfer characteristics of a conventional Ga2O3 MOSFET with a semi-insulating Ga2O3 substrate and Figure 3b shows those of a Ga2O3-on-SiC MOSFET. No significant difference in the threshold voltage between the two MOSFETs can be found, indicating that the transfer characteristics of the device were not affected by the change in substrate material. Although there are differences in the work function between Ga2O3 and SiC, the transfer characteristics of MOSFETs are mainly dependent on the gate dielectric and the work function difference between the gate metal and the channel layer.
Furthermore, the output characteristics of the MOSFETs were discussed. The output characteristics of the two devices before and after considering the thermal effects are shown in Figure 4. The output characteristic of the conventional Ga2O3 MOSFET is shown in Figure 4a, where it is seen that with the increase in gate voltage, the drain current decreases significantly after considering the self-heating effects. The output characteristics of a Ga2O3-on-SiC MOSFET are shown in Figure 4b. It can be observed that the drain current nearly remains identical before and after adding the thermal model. To explore the physical mechanisms, the lattice temperature distribution is calculated and shown in Figure 5. The lattice temperature in the channel region of the Ga2O3-on-SiC MOSFET is only 58 °C, while the temperature in the channel region of the conventional Ga2O3 MOSFET increases to 151 °C. The poor thermal conductivity of Ga2O3 results in a higher lattice temperature, and thus drain current is reduced. The mobility of carriers decreases exponentially with increasing temperature, and according to Equation (8), the conventional Ga2O3 MOSFET has a 39% reduction in carrier mobility compared to the Ga2O3-on-SiC MOSFET. Similar results were reported in previous work, suggesting SiC hetero-substrate is a feasible method for reducing the self-heating effect [18,19,24,25].

4. Breakdown Characteristics

4.1. SiC Thickness and Doping Concentration

Unlike devices consisting of a single semiconductor material, not only the breakdown in the Ga2O3 layer but also the breakdown in the SiC layer shall be considered in Ga2O3-on-SiC MOSFETs. Although the doping concentration of the SiC layer is lower than that of the Ga2O3 layer in the design, the coupling effect can result in an electric field peak at the interface of the two materials. As mentioned previously, the critical breakdown field of SiC is much smaller than that of Ga2O3, so breakdown may happen in SiC rather than Ga2O3. To verify that, the relationship between the breakdown voltage and the doping concentration of the SiC layer was first studied. After that, the impact of the SiC thickness on breakdown voltage was investigated. Since the Ga2O3 channel has n-type doping, only p-type or semi-insulating SiC substrates can be selected to prevent leakage current. For an n-type drift region, a p-type layer is expected to form the reduced surface electric field (RESURF) effect, which has been applied in Si-based and GaN-based MOSFETs and exhibits a satisfying effect on improving the breakdown voltage [42,43,44]. The basic principle is to reduce the peak electric field in the drift region and increase the depletion region width, thereby improving the breakdown voltage of the device.
The relationship between the breakdown voltage and the SiC doping concentration is shown in Figure 6. For comparison, the electric field profile of the conventional Ga2O3 MOSFET at breakdown is illustrated in Figure 6a. A breakdown voltage of 1397 V was achieved. The peak electric field in SiC as a function of the drain voltage was shown in Figure 6b, and the corresponding peak electric field in Ga2O3 was shown in Figure 6c. Note that the breakdown voltage is assumed to correspond to the drain voltage at which the electric field peak in Ga2O3 or SiC reaches a critical field strength. As shown in Figure 6b, the peak electric field in SiC increased as the drain voltage increased, and finally, 3.2 MV/cm was achieved. In Figure 6c, the peak electric field in Ga2O3 was far lower than the critical breakdown strength of Ga2O3 (8 MV/cm), indicating a premature SiC breakdown was formed. From Figure 6b, a breakdown voltage of 795 V was obtained with a doping concentration of 1016 cm−3, and a breakdown voltage of 1057 V was achieved with a doping concentration of 1012 cm−3. As the SiC concentration decreases, the breakdown voltage of the MOSFET increases. With a lower doping concentration, the depletion region width in SiC was increased, thereby allowing a greater drain voltage to be withstood. The peak electric field in the SiC substrate can be effectively suppressed at a low doping concentration, leading to premature breakdown.
As shown in Figure 6c, a wider depletion region in Ga2O3 was formed with the help of the p-type SiC substrate, and charges in the depletion region were redistributed. As a result, the average electric field intensity in Ga2O3 is lowered, and the peak electric field intensity is reduced. Although the peak electric field in Ga2O3 decreases with the increase in p-type doping concentration in SiC, this is at the expense of a higher peak electric field in SiC. By increasing the doping concentration of SiC, its depletion region is reduced, resulting in an earlier breakdown. As can be seen from Figure 6c, the peak electric field in Ga2O3 is always far below its critical breakdown field. Therefore, the breakdown occurs at the interface of SiC/Ga2O3. In other words, a lower doping concentration of SiC is favorable.
According to theoretical calculations, when the doping concentration in SiC is less than 1012 cm−3, the SiC substrate has already achieved semi-insulation. Therefore, for the Ga2O3-on-SiC MOSFET, the semi-insulating SiC substrate can achieve the maximum breakdown voltage. Although the p-type SiC substrate has a RESURF effect on the Ga2O3 layer, it will cause premature breakdown in the substrate, and will not improve the breakdown performance of the device.
In addition to the substrate doping concentration, according to the RESURF theory [45], the thickness of the SiC also has a great impact on the breakdown voltage. Figure 7 shows the peak electric field in SiC and Ga2O3 as a function of SiC layer thickness. As shown in Figure 7a, with the increase in substrate thickness, the breakdown voltage also increases. However, with a further increase in thickness, i.e., thickness ≥ 20 μm, the breakdown voltage of the device tends to saturate. Figure 7b shows the peak electric field in Ga2O3 at various drain voltages. The electric field in Ga2O3 is higher than that in SiC, yet it is still far below its breakdown limit. As a result, the breakdown voltage of the Ga2O3-on-SiC MOSFET is determined to be 1057 V with a thick semi-insulating SiC layer.
To further understand the breakdown mechanisms of the Ga2O3-on-SiC MOSFET, the electric field distribution at the breakdown is shown in Figure 8. The two-dimensional electric field distribution of the Ga2O3-on-SiC MOSFET is shown in Figure 8a, where the thickness and the doping concentration of SiC are 30 μm and 1012 cm−3, respectively. It is seen that two electric field peaks at the gate edge and the drain edge are formed due to the full depletion. In addition, electric field peaks can be found both in the Ga2O3 layer and the SiC layer, which result from the coupling effect [46]. Specifically, the electric field profiles in the Ga2O3 and in the SiC extracted from Figure 8a are shown in Figure 8b. In this case, the intensity of two electric field peaks in the SiC is similar.
Yet for non-ideal conditions, i.e., higher doping concentration and lower thickness of SiC, a visible difference between the electric field peaks can be observed, indicating a lower average electric field in SiC and a reduced breakdown voltage. The electric field profiles in SiC and Ga2O3 with a SiC doping concentration of 1015 cm−3 and a SiC thickness of 30 μm were shown in Figure 8c. Compared with the electric field profile in Figure 8b, the electric field intensity in SiC is lower. As discussed previously, with a higher SiC doping concentration, depletion is more advanced, and the peak electric field is thus reduced. Similarly, we presented the electric field distribution with a 5 μm-thick semi-insulating SiC, as shown in Figure 8d. The profiles of the surface electric fields of SiC and Ga2O3 are shown in Figure 8d. Identical electric field peaks can be found in Ga2O3, yet the electric field in SiC is lower in comparison with the optimal case in Figure 8b. It is well known that the breakdown voltage can be obtained by integrating an electric field profile. Therefore, the dissimilar electric field peaks in SiC are the main reason for the lower breakdown voltage.

4.2. Epitaxy Thickness and Doping

The thickness and doping concentration of the Ga2O3 epitaxial layer also have a significant impact on the breakdown voltage of the device. Therefore, we used the aforementioned method to further investigate the optimal parameters of the Ga2O3 epitaxial layer in this section. A semi-insulating SiC layer was adopted according to the optimization results. In addition, the doping concentration in Ga2O3 varied from 1 × 1016 to 5 × 1017 cm−3, and different thicknesses of Ga2O3 were used in the range from 0.1 μm to 0.5 μm.
In Figure 9a, we exhibit the breakdown voltage as a function of the Ga2O3 doping concentration. The device achieved the highest breakdown voltage of 1057 V at a concentration of 1 × 1017 cm−3. When the doping concentration is less than 1 × 1017 cm−3, the breakdown voltage slowly increases with a higher doping concentration. However, when the doping concentration is greater than 1 × 1017 cm−3, the breakdown voltage of the device decreases sharply. Figure 9b,c shows the electric field distribution profile of the device at a Ga2O3 doping concentration of 1 × 1016 cm−3 and 2 × 1017 cm−3, respectively. At a lower concentration, i.e., 1 × 1016 cm−3, there are two peak electric fields, located below the gate and the drain. Due to the low doping concentration, the Ga2O3 layer was fully depleted, and the electric field intensity close to the drain region was increased. In a conventional Ga2O3 MOSFET, such a low doping concentration is desired since the electric field peak under the gate electrode is reduced, and thus breakdown voltage can be improved. However, for the Ga2O3-on-SiC MOSFET, the electric field distribution in SiC is more vital. It is worth noting that a significant premature breakdown is presented for a lower Ga2O3 doping concentration, and two electric field peaks in SiC were not identical.
As mentioned, the electric field in SiC is coupled with that in Ga2O3. Therefore, a peak electric field in SiC under the drain can be observed, and it has reached the critical breakdown field. When the concentration is high, i.e., 2 × 1017 cm−3, a partial depletion is introduced in Ga2O3, leading to the single electric field peak in the SiC layer. It is well known that partial depletion is harmful to the breakdown voltage, which is the main cause of the rapidly decreased breakdown voltage with a high doping concentration of Ga2O3.
Finally, the relationship between the thickness of the Ga2O3 epitaxial layer and the breakdown voltage was investigated. As shown in Figure 10a, the device achieved a maximum breakdown voltage of 1057 V at an epitaxial thickness of 0.235 μm, and the breakdown voltage increased as the thickness of the epitaxial layer rose until it exceeded 0.235 μm, after which the breakdown voltage began to decrease. Figure 10b,c shows the electric field profiles of the Ga2O3-on-SiC MOSFET with epitaxial thicknesses of 0.1 μm and 0.4 μm, respectively. With a thin epitaxy layer, there are two electric field peaks located under the gate and drain. In Ga2O3, a full depletion region was formed. Less carriers are expected in a thin epitaxy layer, and thus it can be fully depleted by the applied drain voltage. Similarly, the two-peaks property is also validated in the SiC layer, which is the main reason for the higher electric peak near the drain side. Consequently, a lower breakdown voltage is obtained. With an optimal thickness, two similar electric field peaks should be introduced. For a thicker Ga2O3 epitaxial layer, full depletion is hard to perform, and thus the single electric field at the gate edge is observed in Figure 10c. Therefore, we can conclude that the best combination for the Ga2O3 layer is a doping concentration of 1 × 1017 cm−3 and a thickness of 0.235 μm.
One can learn from the above results that the thermal benefit obtained from the Ga2O3-on-SiC MOSFET is verified. However, an early breakdown could be a major hindrance to power devices with a high breakdown voltage. We must admit that the optimal parameters could be varied with different device architectures, such as trench MOSFETs or field-plate MOSFETs. More importantly, to enhance the breakdown voltage of Ga2O3-on-SiC MOSFETs, structural engineering can be considered, including the incorporation of field plate techniques, passivation optimization, variation of lateral doping (VLD), and improvement of epitaxial layer quality. These approaches collectively contribute to improving the breakdown characteristics and overall performance of the MOSFETs [47]. However, the major contribution of this work is to identify the existence of premature breakdown arising from the hetero-material with a lower breakdown strength. Our results suggest that a dedicated device design is required to alleviate premature breakdown.
In addition to the device’s structural optimization, an alternative approach is to introduce a substrate with high breakdown strength as well as high thermal conductivity, such as AlN. Previous reports have indicated the possibility of heteroepitaxy for AlN/Ga2O3 [48]. It is also worth noting that the near junction method [8] by using high thermal conductivity passivation, including amorphous diamond, could be effective for lateral power devices.
It should be noted that except for the above discussion regarding the premature breakdown of SiC, the non-ideal interface is another issue that may lead to premature breakdown. When hetero-epitaxy is conducted, lattice mismatch can lead to a non-ideal interface, resulting in the presence of interface traps or dislocations. Under high electric field conditions, interface traps may undermine carrier mobility [49,50]. Furthermore, dislocations introduce structural distortions, leading to localized strain and an electric field peak, thereby affecting the device’s performance and reliability. It is essential to assess these interface issues to comprehend material properties, electrical characteristics, and breakdown behavior under high electric field conditions [51,52]. Thereby, further investigation is much needed to provide the technical basis for product applications of Ga2O3-based power MOSFETs.

5. Conclusions

In this study, the Ga2O3-on-SiC MOSFET was designed and simulated, including its transfer, output, and breakdown characteristics. For on-state performance, a Ga2O3-on-SiC MOSFET was found to reduce the lattice temperature by nearly 100 °C and increase the drain current by 95%. More importantly, the premature breakdown caused by SiC was studied in this work. By optimizing the structure parameters, including doping concentration and epitaxy thickness, the early breakdown can be well relieved. Compared with p-type SiC, semi-insulating SiC can effectively alleviate SiC premature breakdown, obtaining a higher breakdown voltage. Additionally, a thick (>20 μm) semi-insulating SiC substrate is beneficial to obtain a higher breakdown voltage. As for the Ga2O3 epitaxial layer, under the conditions of a concentration of 1 × 1017 cm−3 and a thickness of 0.235 μm, a breakdown voltage of 1057 V is achieved, which is 75% of the breakdown voltage for the conventional Ga2O3 MOSFET. In summary, by optimizing the device structure, the Ga2O3-on-SiC MOSFET can realize superior thermal management performance without sacrificing a significant breakdown voltage.

Author Contributions

Conceptualization, M.Z. and L.W.; Data curation, M.Z. and L.W.; Formal analysis, M.Z. and L.W.; Methodology, W.T. and Y.G.; Software, L.W.; Validation, K.Y. and J.Y.; Writing—original draft, M.Z. and L.W.; Writing—review and editing, W.T. and Y.G. All authors have read and agreed to the published version of the manuscript.

Funding

The National Key R&D Program of China (Grant No. 2022YFB3605404), the China Postdoctoral Science Foundation (Grant No. 2022M721689), and the National Natural Science Foundation of China (Grant No. 61874059).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Irmscher, K.; Galazka, Z.; Pietsch, M.; Uecker, R.; Fornari, R. Electrical properties of β-Ga2O3 single crystals grown by the Czochralski method. J. Appl. Phys. 2011, 110, 063720. [Google Scholar] [CrossRef]
  2. Fujita, S.; Oda, M.; Kaneko, K.; Hitora, T. Evolution of corundum-structured III-oxide semiconductors: Growth, properties, and devices. Jpn. J. Appl. Phys. 2016, 55, 1202A3. [Google Scholar] [CrossRef] [Green Version]
  3. Sugiura, T.; Nakano, N. Impact Ionization and Critical Electric Field in ⟨010⟩-Oriented β-Ga2O3 Schottky Barrier Diode. IEEE Trans. Electron Devices 2022, 69, 3068–3072. [Google Scholar] [CrossRef]
  4. Kamimura, T.; Nakata, Y.; Wong, M.H.; Higashiwaki, M. Normally-Off Ga2O3 MOSFETs With Unintentionally Nitrogen-Doped Channel Layer Grown by Plasma-Assisted Molecular Beam Epitaxy. IEEE Electron. Device Lett. 2019, 40, 1064–1067. [Google Scholar] [CrossRef]
  5. Higashiwaki, M.; Sasaki, K.; Murakami, H.; Kumagai, Y.; Koukitu, A.; Kuramata, A.; Masui, T.; Yamakoshi, S. Recent progress in Ga2O3 power devices. Semicond. Sci. Technol. 2016, 31, 034001. [Google Scholar] [CrossRef]
  6. Wong, M.H.; Higashiwaki, M. Vertical β-Ga2O3 Power Transistors: A Review. IEEE Trans. Electron Devices 2020, 67, 3925–3937. [Google Scholar] [CrossRef]
  7. Wong, M.H.; Morikawa, Y.; Sasaki, K.; Kuramata, A.; Yamakoshi, S.; Higashiwaki, M. Characterization of channel temperature in Ga2O3 metal-oxide-semiconductor field-effect transistors by electrical measurements and thermal modeling. Appl. Phys. Lett. 2016, 109, 193503. [Google Scholar] [CrossRef]
  8. Chatterjee, B.; Zeng, K.; Nordquist, C.D.; Singisetti, U.; Choi, S. Device-Level Thermal Management of Gallium Oxide Field-Effect Transistors. IEEE Trans. Compon. Packag. Manuf. Technol. 2019, 9, 2352–2365. [Google Scholar] [CrossRef]
  9. Wang, B.; Xiao, M.; Zhang, Z.; Wang, Y.; Qin, Y.; Song, Q.; Lu, G.-Q.; Ngo, K.; Zhang, Y. Chip Size Minimization for Wide and Ultrawide Bandgap Power Devices. IEEE Trans. Electron Devices 2023, 70, 633–639. [Google Scholar] [CrossRef]
  10. Slomski, M.; Blumenschein, N.; Paskov, P.P.; Muth, J.F.; Paskova, T. Anisotropic thermal conductivity of β-Ga2O3 at elevated temperatures: Effect of Sn and Fe dopants. J. Appl. Phys. 2017, 121, 235104. [Google Scholar] [CrossRef]
  11. Zhu, L.; Li, Z.; Hou, K. Effect of radical scavenger on electrical tree in cross-linked polyethylene with large harmonic superimposed DC voltage. High Volt. 2022, 1–10. [Google Scholar] [CrossRef]
  12. Yang, S.T.; Li, X.Y.; Yu, T.L.; Wang, J.; Fang, H.; Nie, F.; He, B.; Zhao, L.; Lü, W.M.; Yan, S.S.; et al. High-Performance Neuromorphic Computing Based on Ferroelectric Synapses with Excellent Conductance Linearity and Symmetry. Adv. Funct. Mater. 2022, 32, 2202366. [Google Scholar] [CrossRef]
  13. Yang, L.; Huang, H.; Xi, Z.; Zheng, L.; Xu, S.; Tian, G.; Zhai, Y.; Guo, F.; Kong, L.; Wang, Y.; et al. Simultaneously achieving giant piezoelectricity and record coercive field enhancement in relaxor-based ferroelectric crystals. Nat. Commun. 2022, 13, 2444. [Google Scholar] [CrossRef]
  14. Yan, A.; Xiang, J.; Cao, A.; He, Z.; Cui, J.; Ni, T.; Huang, Z.; Wen, X.; Girard, P. Quadruple and Sextuple Cross-Coupled SRAM Cell Designs With Optimized Overhead for Reliable Applications. IEEE Trans. Device Mater. Reliab. 2022, 22, 282–295. [Google Scholar] [CrossRef]
  15. Yan, A.; Li, Z.; Cui, J.; Huang, Z.; Ni, T.; Girard, P.; Wen, X. Designs of Two Quadruple-Node-Upset Self-Recoverable Latches for Highly Robust Computing in Harsh Radiation Environments. IEEE Trans. Aerosp. Electron. Syst. 2022, 1–13. [Google Scholar] [CrossRef]
  16. Xiang, J.; Yang, W.; Liao, H.; Li, P.; Chen, Z.; Huang, J. Design and thermal performance of thermal diode based on the asymmetric flow resistance in vapor channel. Int. J. Therm. Sci. 2023, 191, 108345. [Google Scholar] [CrossRef]
  17. Nepal, N.; Katzer, D.S.; Downey, B.P.; Wheeler, V.D.; Nyakiti, L.O.; Storm, D.F.; Hardy, M.T.; Freitas, J.A.; Jin, E.N.; Vaca, D.; et al. Heteroepitaxial growth of β-Ga2O3 films on SiC via molecular beam epitaxy. J. Vac. Sci. Technol. A 2020, 38, 063406. [Google Scholar] [CrossRef]
  18. Russell, S.A.O.; Perez-Tomas, A.; McConville, C.F.; Fisher, C.A.; Hamilton, D.P.; Mawby, P.A.; Jennings, M.R. Heteroepitaxial Beta-Ga2O3 on 4H-SiC for an FET With Reduced Self Heating. IEEE J. Electron Devices Soc. 2017, 5, 256–261. [Google Scholar] [CrossRef]
  19. Lee, D.; Kim, H.W.; Kim, J.; Moon, J.H.; Lee, G.; Kim, J. Ultra-Wide Bandgap β-Ga2O3 Heterojunction Field-Effect Transistor Using p-Type 4H-SiC Gate for Efficient Thermal Management. ECS J. Solid State Sci. Technol. 2020, 9, 065006. [Google Scholar] [CrossRef]
  20. Hatakeyama, T.; Shinohe, T. Reverse Characteristics of a 4H-SiC Schottky Barrier Diode. Mater. Sci. Forum 2002, 389–393, 1169–1172. [Google Scholar] [CrossRef]
  21. Chen, J.; Zhang, Z.; Lu, H. Structure design and properties investigation of Bi2O2Se/graphene van der Waals heterojunction from first-principles study. Surf. Interfaces 2022, 33, 102289. [Google Scholar] [CrossRef]
  22. Li, S.; Chen, J.; He, X.; Zheng, Y.; Yu, C.; Lu, H. Comparative study of the micro-mechanism of charge redistribution at metal-semiconductor and semimetal-semiconductor interfaces: Pt(Ni)-MoS2 and Bi-MoS2(WSe2) as the prototype. Appl. Surf. Sci. 2023, 623, 157036. [Google Scholar] [CrossRef]
  23. Li, S.; Yang, L.; Liu, Z.; Zhang, M.; Guo, Y.; Tang, W. Hybrid PEDOT:PSS/SiC heterojunction UV photodetector with superior self-powered responsivity over A/W level. Appl. Phys. Lett. 2023, 122, 191102. [Google Scholar] [CrossRef]
  24. Wang, Y.; Xu, W.; Han, G.; You, T.; Mu, F.; Hu, H.; Liu, Y.; Zhang, X.; Huang, H.; Suga, T.; et al. Channel Properties of Ga2O3-on-SiC MOSFETs. IEEE Trans. Electron Devices 2021, 68, 1185–1189. [Google Scholar] [CrossRef]
  25. Shin, M.C.; Lee, Y.J.; Kim, D.H.; Jung, S.W.; Schweitz, M.A.; Shin, W.H.; Oh, J.M.; Park, C.; Koo, S.M. Improved Electrical Characteristics of Gallium Oxide/P-Epi Silicon Carbide Static Induction Transistors with UV/Ozone Treatment Fabricated by RF Sputter. Materials 2021, 14, 1296. [Google Scholar] [CrossRef] [PubMed]
  26. Zhou, X.; Liu, Q.; Xu, G.; Zhou, K.; Xiang, X.; He, Q.; Hao, W.; Jian, G.; Zhao, X.; Long, S. Realizing High-Performance β-Ga2O3 MOSFET by Using Variation of Lateral Doping: A TCAD Study. IEEE Trans. Electron Devices 2021, 68, 1501–1506. [Google Scholar] [CrossRef]
  27. Zhao, Y. Co-precipitated Ni/Mn shell coated nano Cu-rich core structure: A phase-field study. J. Mater. Res. Technol. 2022, 21, 546–560. [Google Scholar] [CrossRef]
  28. Zhang, L.; Xiong, D.; Su, Z.; Li, J.; Yin, L.; Yao, Z.; Wang, G.; Zhang, L.; Zhang, H. Molecular dynamics simulation and experimental study of tin growth in SAC lead-free microsolder joints under thermo-mechanical-electrical coupling. Mater. Today Commun. 2022, 33, 104301. [Google Scholar] [CrossRef]
  29. Wang, D.; Xu, K.D.; Luo, S.; Cui, Y.; Zhang, L.; Cui, J. A high Q-factor dual-band terahertz metamaterial absorber and its sensing characteristics. Nanoscale 2023, 15, 3398–3407. [Google Scholar] [CrossRef]
  30. Muhammad, I.; Ali, A.; Zhou, L.; Zhang, W.; Wong, P.K.J. Vacancy-engineered half-metallicity and magnetic anisotropy in CrSI semiconductor monolayer. J. Alloys Compd. 2022, 909, 164797. [Google Scholar] [CrossRef]
  31. Luo, Z.-Z.; Cai, S.; Hao, S.; Bailey, T.P.; Luo, Y.; Luo, W.; Yu, Y.; Uher, C.; Wolverton, C.; Dravid, V.P.; et al. Extraordinary role of Zn in enhancing thermoelectric performance of Ga-doped n-type PbTe. Energy Environ. Sci. 2022, 15, 368–375. [Google Scholar] [CrossRef]
  32. Li, Q.; Yuan, L.; Zhang, F.; Li, H.; Xiao, G.; Chen, Y.; Sun, T.; Liu, X.; Fu, T. Novel SiC/Si heterojunction LDMOS with electric field modulation effect by reversed L-shaped field plate. Results Phys. 2020, 16, 102837. [Google Scholar] [CrossRef]
  33. Torky, M.; Chow, T.P. Determination of Effective Critical Breakdown Field in 4H-SiC Superjunction Devices. Mater. Sci. Forum 2022, 1062, 560–564. [Google Scholar] [CrossRef]
  34. Lu, A.; Elwailly, A.; Zhang, Y.; Wong, H.Y. Study of Vertical Ga2O3 FinFET Short Circuit Ruggedness using Robust TCAD Simulation. ECS J. Solid State Sci. Technol. 2022, 11, 115001. [Google Scholar] [CrossRef]
  35. Wang, L.; Zhang, M.; Liu, Z.; Yang, L.; Li, S.; Tang, W.; Guo, Y. Investigation of Lateral Gallium Oxide MOSFET with Discrete Field Plate by Calibrated Numerical Simulation. In Proceedings of the 2022 19th China International Forum on Solid State Lighting & 2022 8th International Forum on Wide Bandgap Semiconductors (SSLCHINA: IFWS), Suzhou, China, 7–10 February 2023; pp. 181–184. [Google Scholar]
  36. Park, J.; Hong, S.-M. Simulation Study of Enhancement Mode Multi-Gate Vertical Gallium Oxide MOSFETs. ECS J. Solid State Sci. Technol. 2019, 8, Q3116–Q3121. [Google Scholar] [CrossRef]
  37. Arora, N.D.; Hauser, J.R.; Roulston, D.J. Electron and hole mobilities in silicon as a function of concentration and temperature. IEEE Trans. Electron Devices 1982, 29, 292–295. [Google Scholar] [CrossRef]
  38. Moser, N.; McCandless, J.; Crespo, A.; Leedy, K.; Green, A.; Neal, A.; Mou, S.; Ahmadi, E.; Speck, J.; Chabak, K.; et al. Ge-Doped β-Ga2O3 MOSFETs. IEEE Electron Device Lett. 2017, 38, 775–778. [Google Scholar] [CrossRef]
  39. Ma, N.; Tanen, N.; Verma, A.; Guo, Z.; Luo, T.; Xing, H.G.; Jena, D. Intrinsic electron mobility limits in β-Ga2O3. Appl. Phys. Lett. 2016, 109, 212101. [Google Scholar] [CrossRef] [Green Version]
  40. Caughey, D.M.; Thomas, R. Carrier mobilities in silicon empirically related to doping and field. Proc. IEEE 1967, 55, 2192–2193. [Google Scholar] [CrossRef]
  41. Jia, X.; Hu, H.; Han, G.; Liu, Y.; Hao, Y. Analytical Model for the Channel Maximum Temperature in Ga2O3 MOSFETs. Nanoscale Res. Lett. 2021, 16, 29. [Google Scholar] [CrossRef]
  42. Yang, K.; Guo, Y.; Zhang, J.; Yao, J.; Li, M.; Du, L.; Huang, X. An Analytical Breakdown Model for the SOI LDMOS With Arbitrary Drift Doping Profile by Using Effective Substrate Voltage Method. IEEE J. Electron Devices Soc. 2020, 8, 49–56. [Google Scholar] [CrossRef]
  43. Wu, H.; Duan, B.-X.; Yang, L.-Y.; Yang, Y.-T. Theoretical analytic model for RESURF AlGaN/GaN HEMTs. Chin. Phys. B 2019, 28, 027302. [Google Scholar] [CrossRef]
  44. Zhang, J.; Guo, Y.-F.; Yang, K.-M.; Huang, C.-Y.; Hu, F.-R. Modeling of the Variation of Lateral Doping (VLD) Lateral Power Devices via 1-D Analysis Using Effective Concentration Profile Concept. IEEE J. Electron Devices Soc. 2019, 7, 990–996. [Google Scholar] [CrossRef]
  45. Yufeng, G.; Zhigong, W.; Gene, S. A three-dimensional breakdown model of SOI lateral power transistors with a circular layout. J. Semicond. 2009, 30, 114006. [Google Scholar] [CrossRef]
  46. Qiao, M.; Li, Y.; Zhou, X.; Li, Z.; Zhang, B. A 700-V Junction-Isolated Triple RESURF LDMOS with N-Type Top Layer. IEEE Electron Device Lett. 2014, 35, 774–776. [Google Scholar] [CrossRef]
  47. Zhang, M.; Liu, Z.; Yang, L.; Yao, J.; Chen, J.; Zhang, J.; Wei, W.; Guo, Y.; Tang, W. β-Ga2O3-Based Power Devices: A Concise Review. Crystals 2022, 12, 406. [Google Scholar] [CrossRef]
  48. Wang, G.; Zhou, Y. Thermal Management Modeling for β-Ga2O3-Highly Thermal Conductive Substrates Heterostructures. IEEE Trans. Compon. Packag. Manuf. Technol. 2022, 12, 638–646. [Google Scholar] [CrossRef]
  49. Xia, X.; Li, J.-S.; Chiang, C.-C.; Ren, F.; Pearton, S. Fabrication and Device Performance of 2.7 Kv/2.5 A NiO/Ga2O3 Heterojunction Power Rectifiers. ECS Trans. 2023, 111, 103. [Google Scholar] [CrossRef]
  50. Li, S.; Guo, D.; Li, P.; Wang, X.; Wang, Y.; Yan, Z.; Liu, Z.; Zhi, Y.; Huang, Y.; Wu, Z.; et al. Ultrasensitive, Superhigh Signal-to-Noise Ratio, Self-Powered Solar-Blind Photodetector Based on n-Ga2O3/p-CuSCN Core-Shell Microwire Heterojunction. ACS Appl. Mater. Interfaces 2019, 11, 35105–35114. [Google Scholar] [CrossRef]
  51. Cheng, Z.; Mu, F.; You, T.; Xu, W.; Shi, J.; Liao, M.E.; Wang, Y.; Huynh, K.; Suga, T.; Goorsky, M.S.; et al. Thermal Transport across Ion-Cut Monocrystalline beta-Ga(2)O(3) Thin Films and Bonded beta-Ga2O3-SiC Interfaces. ACS Appl. Mater. Interfaces 2020, 12, 44943–44951. [Google Scholar] [CrossRef]
  52. Wu, F.; Wang, Y.; Jian, G.; Xu, G.; Zhou, X.; Guo, W.; Du, J.; Liu, Q.; Dun, S.; Yu, Z.; et al. Superior Performance β-Ga2O3 Junction Barrier Schottky Diodes Implementing p-NiO Heterojunction and Beveled Field Plate for Hybrid Cockcroft–Walton Voltage Multiplier. IEEE Trans. Electron Devices 2023, 70, 1199–1205. [Google Scholar] [CrossRef]
Figure 1. Schematic diagrams of the (a) conventional Ga2O3 MOSFET and (b) Ga2O3-on-SiC MOSFET.
Figure 1. Schematic diagrams of the (a) conventional Ga2O3 MOSFET and (b) Ga2O3-on-SiC MOSFET.
Crystals 13 00917 g001
Figure 2. Calibration of simulated transfer curves with experimental results from [38].
Figure 2. Calibration of simulated transfer curves with experimental results from [38].
Crystals 13 00917 g002
Figure 3. Transfer characteristics of (a) conventional Ga2O3 MOSFET and (b) Ga2O3-on-SiC MOSFET.
Figure 3. Transfer characteristics of (a) conventional Ga2O3 MOSFET and (b) Ga2O3-on-SiC MOSFET.
Crystals 13 00917 g003
Figure 4. Output characteristics of (a) conventional Ga2O3 MOSFET and (b) Ga2O3-on-SiC MOSFET. The self-heating effect is considered and compared.
Figure 4. Output characteristics of (a) conventional Ga2O3 MOSFET and (b) Ga2O3-on-SiC MOSFET. The self-heating effect is considered and compared.
Crystals 13 00917 g004
Figure 5. Lattice temperature distribution of (a) the Ga2O3-on-SiC MOSFET and (b) the conventional Ga2O3 MOSFET. (c,d) are enlarged channel regions of Ga2O3-on-SiC MOSFET and conventional Ga2O3 MOSFET with temperature distribution displayed, respectively.
Figure 5. Lattice temperature distribution of (a) the Ga2O3-on-SiC MOSFET and (b) the conventional Ga2O3 MOSFET. (c,d) are enlarged channel regions of Ga2O3-on-SiC MOSFET and conventional Ga2O3 MOSFET with temperature distribution displayed, respectively.
Crystals 13 00917 g005
Figure 6. (a) Electric field distribution of the conventional Ga2O3 MOSFET at the breakdown. A breakdown voltage of 1397 V is obtained. Peak electric fields in (b) SiC and (c) Ga2O3 as a function of drain voltage with various SiC doping concentrations.
Figure 6. (a) Electric field distribution of the conventional Ga2O3 MOSFET at the breakdown. A breakdown voltage of 1397 V is obtained. Peak electric fields in (b) SiC and (c) Ga2O3 as a function of drain voltage with various SiC doping concentrations.
Crystals 13 00917 g006
Figure 7. Peak electric field in (a) SiC and (b) Ga2O3 as a function of drain voltage with different SiC thicknesses.
Figure 7. Peak electric field in (a) SiC and (b) Ga2O3 as a function of drain voltage with different SiC thicknesses.
Crystals 13 00917 g007
Figure 8. The electric field distribution in a Ga2O3-on-SiC MOSFET. (a) Two-dimensional distribution of the device with a 30 μm semi-insulating SiC. The electric field profiles in (b,c) for Ga2O3 and SiC are extracted from the red cutline and the yellow cutline, respectively. Electric field profiles in Ga2O3 and SiC with (b) 30 μm semi-insulating SiC, (c) 30 μm SiC with a doping concentration of 1015 cm−3 and (d) 5 μm semi-insulating SiC.
Figure 8. The electric field distribution in a Ga2O3-on-SiC MOSFET. (a) Two-dimensional distribution of the device with a 30 μm semi-insulating SiC. The electric field profiles in (b,c) for Ga2O3 and SiC are extracted from the red cutline and the yellow cutline, respectively. Electric field profiles in Ga2O3 and SiC with (b) 30 μm semi-insulating SiC, (c) 30 μm SiC with a doping concentration of 1015 cm−3 and (d) 5 μm semi-insulating SiC.
Crystals 13 00917 g008
Figure 9. (a) The relationship between breakdown voltage and doping concentration of the Ga2O3 epitaxial layer. The electric field profiles at the breakdown with a doping concentration of (b) 1 × 1016 cm−3 and (c) 2 × 1017 cm−3.
Figure 9. (a) The relationship between breakdown voltage and doping concentration of the Ga2O3 epitaxial layer. The electric field profiles at the breakdown with a doping concentration of (b) 1 × 1016 cm−3 and (c) 2 × 1017 cm−3.
Crystals 13 00917 g009
Figure 10. (a) The relationship between breakdown voltage and thickness of the Ga2O3 epitaxial layer. Electric field distributions at the breakdown with an epitaxial thickness of (b) 0.1 μm and (c) 0.4 μm.
Figure 10. (a) The relationship between breakdown voltage and thickness of the Ga2O3 epitaxial layer. Electric field distributions at the breakdown with an epitaxial thickness of (b) 0.1 μm and (c) 0.4 μm.
Crystals 13 00917 g010
Table 1. Device structural parameters.
Table 1. Device structural parameters.
Structural ParametersUnitValue
Substrate thickness tsubμm200
Drain to gate distance Lgdμm7.5
Al2O3 thickness toxnm20
Gate work function WF eV5.3
Substrate temperature TsubK300
Ga2O3 channel thickness tchμm0.235 for on-state, 0.1–0.5 for off-state
Ga2O3 channel doping Dchcm−31 × 1017 for on-state, 1 × 1016–5 × 1017 for off-state
SiC layer thickness tSiCμm30 for on-state, 5–30 for off-state
SiC layer doping DSiCcm−31 × 1012 for on-state, 1 × 1015–1 × 1016 for off-state
Table 2. Arora model variables in Ga2O3.
Table 2. Arora model variables in Ga2O3.
VariablesElectronHoleUnit
Amin130.016cm2/V∙s
αm−0.57−0.571
Ad2350.2cm2/V∙s
αd0.780.781
AN1.1 × 10181.25 × 1017cm−3
αN2.42.41
Aa0.780.781
αa−0.146−0.1461
Table 3. Ga2O3 Canali model variables.
Table 3. Ga2O3 Canali model variables.
VariablesValueUnit
α01
β011
βexp01
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, M.; Wang, L.; Yang, K.; Yao, J.; Tang, W.; Guo, Y. Breakdown Characteristics of Ga2O3-on-SiC Metal-Oxide-Semiconductor Field-Effect Transistors. Crystals 2023, 13, 917. https://doi.org/10.3390/cryst13060917

AMA Style

Zhang M, Wang L, Yang K, Yao J, Tang W, Guo Y. Breakdown Characteristics of Ga2O3-on-SiC Metal-Oxide-Semiconductor Field-Effect Transistors. Crystals. 2023; 13(6):917. https://doi.org/10.3390/cryst13060917

Chicago/Turabian Style

Zhang, Maolin, Lei Wang, Kemeng Yang, Jiafei Yao, Weihua Tang, and Yufeng Guo. 2023. "Breakdown Characteristics of Ga2O3-on-SiC Metal-Oxide-Semiconductor Field-Effect Transistors" Crystals 13, no. 6: 917. https://doi.org/10.3390/cryst13060917

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop