Next Article in Journal
Comparative Study of Sentinel-1-Focused and Simulated SAR Images Using LiDAR Point Cloud Modeling for Coastal Areas
Previous Article in Journal
Interpretable Multi-Channel Capsule Network for Human Motion Recognition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Large-Scale β-Ga2O3 Trench MOS-Type Schottky Barrier Diodes with 1.02 Ideality Factor and 0.72 V Turn-On Voltage

State Key Laboratory of ASIC and System, School of Microelectronics, Zhangjiang Fudan International Innovation Center, Fudan University, Shanghai 200433, China
*
Author to whom correspondence should be addressed.
Electronics 2023, 12(20), 4315; https://doi.org/10.3390/electronics12204315
Submission received: 18 September 2023 / Revised: 14 October 2023 / Accepted: 15 October 2023 / Published: 18 October 2023
(This article belongs to the Section Power Electronics)

Abstract

:
β-Ga2O3 Schottky barrier diodes (SBDs) suffer from the electric field crowding and barrier height lowering effect, resulting in a low breakdown voltage (BV) and high reverse leakage current. Here, we developed β-Ga2O3 trench MOS-type Schottky barrier diodes (TMSBDs) on β-Ga2O3 single-crystal substrates with halide vapor phase epitaxial layers based on ultraviolet lithography and dry etching. The 1 / C 2 V   plots are deflected at 2.24 V, which is caused by the complete depletion in the mesa region of the TMSBDs. A close-to-unity ideality factor of 1.02 and a low turn-on voltage of 0.72 V are obtained. This is due to the low interface trap density in the metal/semiconductor interface of TMSBDs, as confirmed by the current–voltage ( I V ) hysteresis measurements. The specific on-resistance calculated with the actual Schottky contact area increases as the area ratio ( A R ) increases because of the current spreading phenomenon. Furthermore, the reverse leakage current of the TMSBDs is smaller and the BV is increased by 120 V compared with the regular SBD. This work paves the way for further improving the overall performance of β-Ga2O3 TMSBDs.

1. Introduction

Recently, the ultra-wide bandgap semiconductor β-Ga2O3 has attracted extensive attention in power device applications owing to its outstanding material performance [1]. β-Ga2O3 has a wide bandgap of 4.9 eV [2], a high critical electric field of 8 MV/cm [3], and its Baliga’s figure-of-merit (3214) is ~10 or 4 times larger than that of 4H-SiC or GaN [4]. Large-scale β-Ga2O3 wafers can be fabricated by low-cost melt growth methods such as floating zone, edge-defined film-fed growth, Czochralski, etc. [5,6,7]. High-quality thin films can be epitaxially grown on β-Ga2O3 substrates using various methods such as molecular beam epitaxy, metal–organic chemical vapor deposition and halide vapor phase epitaxy [8,9,10,11]. In addition, the chemical stability and thermal stability of β-Ga2O3 are excellent [12,13]. Therefore, power devices such as metal–oxide semiconductor field-effect transistors (MOSFETs) [14,15,16,17] and SBDs [18,19,20,21] fabricated on β-Ga2O3 wafers have shown strong competitiveness.
On the account of the electric field crowding at the Schottky contact edge and barrier height lowering effect [22], there is still a large gap on the figure-of-merit between β-Ga2O3 SBDs and the material limit. For this reason, devices designed with effective electric field management are crucial to power application. Many edge termination methods, such as field plates [18,21], ion implantation [20,23], thermally oxidized termination [24], and heterogeneous termination based p-type NiO [25,26], have been proposed to improve the reverse performance of the device. For field plates, the dielectric is unable to withstand the high-density electric field, resulting in a premature breakdown of the device. The ion implantation and thermally oxidized termination could only realize a weak p-type or high-resistance state, which does not achieve effective electric field management. P-type NiO makes up the gap of the unavailability of p-type doping for β-Ga2O3. However, it is necessary to further enhance the lattice mismatch of heterojunction interfaces and improve the stability of NiO materials. Additionally, these three types of methods have little inhibitory effect on the barrier height lowering effect. To address these issues, the trench MOS structure is introduced here [19,27,28,29,30,31]. Since the semiconductor is depleted at the Schottky contact, the trench MOS structure could eliminate the effect of electric field crowding. The peak of the electric field transfers from the Schottky contact edge to the corner of the trench. Moreover, the depletion region generated by the MOS capacitance on the sidewalls of the device effectively reduces the metal/semiconductor surface field and suppresses the barrier height lowering effect [29,32]. In the meanwhile, the reverse leakage current can be effectively reduced and BV can be further enhanced.
There has been a growing interest in β-Ga2O3 TMSBDs and significant progress has been made in this field. Sasaki et al. prepared β-Ga2O3 TMSBDs via dry etching and chemical–mechanical polishing, and obtained a BV of 240 V [19]. Li et al. utilized field plates and a trench MOS structure for improving the breakdown characteristics, achieved a power figure-of-merit of 0.95 GW/cm2 [29]. Jian et al. have shown that TMSBDs exhibited superior performance and stability compared to regular SBDs under various temperatures [30]. Nevertheless, few studies have explored the forward characteristics of β-Ga2O3 TMSBDs, such as the ideality factor and turn-on voltage. β-Ga2O3 SBDs still face challenges of a relatively high ideality factor (>1.1) and turn-on voltage (>0.8 V) [19,29,30].
In this work, large-scale β-Ga2O3 TMSBDs have been demonstrated using ultraviolet lithography and dry etching. A turn-on voltage of 0.72 V and an ideality factor of 1.02 were achieved, which was attributed to the low interface trap density. Compared with regular SBDs, the BV of the TMSBDs was increased by 120 V and the reverse leakage current reduced by several orders of magnitude.

2. Materials and Methods

Figure 1 shows the cross-sectional schematic of the β-Ga2O3 TMSBDs. The (001) n-type β-Ga2O3 used in this experiment was purchased from Novel Crystal Technology, Japan. The thickness and doping concentration of the Sn-doped β-Ga2O3 substrate are 598 μm and ~ 5.1 × 10 18   c m 3 , respectively. The Si-doped β-Ga2O3 drift layer was grown on the β-Ga2O3 substrate via halide vapor phase epitaxy, and its thickness and doping concentration are 11 μm and 3.2 × 10 16   c m 3 , respectively. The mesa widths are 20, 24, 28, and 32 μm, respectively, and the trench width is fixed at 40 μm. The mesa depth is 1.64 μm and the mesa length is 300 μm.
Figure 2 presents the process flow of the fabricated TMSBDs. First, the β-Ga2O3 wafers were cleaned with acetone, isopropanol, and deionized water in an ultrasonic environment to effectively remove surface impurities. Then, Ni hard mask patterns were formed by ultraviolet lithography, physical vapor deposition, and lift-off process. Next, a periodic structure composing of trenches and mesas was fabricated on a β-Ga2O3 wafer via inductive coupled plasma-reactive ion etching, and the etching gases were BCl3 and Ar. Then, a Ni hard mask was removed using a solution of hydrochloric acid, and the bottom of the trench was smoothed to alleviate the concentration of the electric field [27,33]. The Ti (20 nm)/Au (100 nm) cathode electrode was deposited on the backside of the β-Ga2O3 wafer via electron beam evaporation. The ohmic contact was formed through rapid thermal annealing in N2 ambient at 470 °C for 60 s. Next, a 50 nm Al2O3 film was deposited on the surface of the structure via atomic layer deposition. An overlay technique was used to open a hole at the top of the mesa and a dry etching method was used to remove the exposed Al2O3. Finally, the anode electrode of Ni (50 nm)/Au (100 nm) was deposited on the top of the structure. A regular SBD with a circular anode (a diameter of 300 μm) was fabricated for comparison. The capacitance–voltage ( C V ), forward current density–voltage ( J V ), and reverse J V characteristics were measured using a Keysight 4294A meter, a Keysight B1500A semiconductor analyzer, and a Keysight B1505A high-biased analyzer at room temperature, respectively.
Figure 3 shows the optical micrograph of the 24 μm β-Ga2O3 TMSBD and regular SBD. The area within dashed red lines is the anode area with a size of 320 μm × 640 μm. The area within dashed blue lines is the Schottky contact area of one mesa. The area ratio ( A R ) is defined as the ratio of the actual Schottky contact area to the anode area. The A R s corresponding to TMSBDs with a Wmesa of 20, 24, 28, and 32 μm are 15.8%, 21.1%, 26.4%, and 31.6%, respectively. The TMSBDs fabricated in the [010] orientation had the lowest interface trap density compared to those fabricated in the [100] orientation [28]. Therefore, the trench structures of all devices demonstrated here have the [010] orientation.

3. Results and Discussion

Figure 4a,b shows the C V and 1 / C 2 V characteristics for the regular SBD and TMSBDs measured at 1 MHz. It is seen that the 1 / C 2 V plot of the regular SBD is a standard straight line, while the 1 / C 2 V plots of TMSBDs are deflected at a certain voltage. We linearly fitted the 1 / C 2 V plots before and after deflection of the TMSBDs, and calculated the voltage at the intersection of the fitted plots as 2.24 V. According to the designed device structure, we give the following explanation. Before the deflection point, the capacitances of the TMSBDs are mainly composed of the Schottky junction capacitance and the MOS capacitances at the bottom of the trench, the side wall of the mesa, and the top of the mesa. As the voltage increases, the depletion region in the mesa keeps expanding. At the deflection point, the mesa is just completely depleted. After the deflection point, the depletion region extends to the semiconductor region below the mesa, and only the MOS capacitances at the bottom and top of the trench are preserved [19]. As a result, the falling rate of capacitance decreases with the decrease in voltage and the 1 / C 2 V plots are deflected.
The capacitance of an SBD is given as follows [34]:
C = q ε 0 ε n N D 2 ( V b i V r ) 1 / 2
where q is the electric charge, N D is the donor concentration, V b i is the built-in potential,   V r is the applied voltage, ε 0 is the vacuum dielectric constant, and ε n = 10 is the relative dielectric constant of β-Ga2O3 [35]. The N D was determined to be 1.49 × 10 16   c m 3 based on the 1 / C 2 V plot. The V b i was 0.95   V from extrapolation to 1 / C 2 = 0 . Taking an electron effective mass of β-Ga2O3  m n * = 0.342 m 0 ( m 0 is the electron rest mass), the local conduction band density of states ( N c ) of 5.02 × 10 18   c m 3 and the energy difference between the conduction band edge and the Fermi energy ( E C E F ) of 0.15   e V can be obtained [36]. The corresponding Schottky barrier height was q ϕ B = q V b i + E C E F = 1.1   e V .
The depletion layer width ( W ) of the Schottky contact at the deflection point voltage of 1 / C 2 V plots was 1.63 μm using the following equation:
W = 2 ε n ( V b i V ) q N D 1 / 2
W is 100 nm lower than the mesa depth, consistent with the theoretical analysis above. The depletion of the mesa region is attributed to the depletion of the MOS structure in the sidewall and Schottky contact structure in the top of the devices. Owing to the slower depletion rate of the MOS structure in the sidewall compared to that of Schottky contact structure in the top of the devices, the difference between W and mesa depth will decrease with an increasing width-to-depth ratio of the trench.
Figure 5a,b represents the frequency dependences of C V and 1 / C 2 V curves for a 28 μm TMSBD at room temperature. It has been observed that capacitance decreases with increasing frequency. This could be caused by interface traps resulting from material defects, lattice mismatches, and etching defects [37]. At a low frequency, interface trap charges will capture and emit electrons, contributing additional capacitance. The capacitance gradually decreases as the frequency increases from 1 kHz to 100 kHz, while it hardly changes after 800 kHz. This is due to the slow majority of traps near the conduction band edge which cannot follow the AC signal, therefore its capacitance is closer to the ideal value [35,38].
Figure 6a depicts the forward J V characteristics of the TMSBDs in linear scales. As the A R increases, the forward current density of the TMSBDs increases and the differential specific on-resistance ( R o n , s p ) calculated with the anode area decreases, which is attributed to the restricted current conduction path [30]. On the one hand, different A R s correspond to different current conduction paths. On the other hand, negative charges at the fin sidewalls will cause depletion of the sidewall, thereby reducing the current density of the TMSBDs. The turn-on voltage of the TMSBDs is about 0.72 V taking the voltage at the forward current of 100   m A / c m 2 . It has no obvious relationship with the W m e s a or A R . The minimum R o n , s p of 24.15   m Ω c m 2 is measured on a TMSBD with a W m e s a of 32 μm, while the maximum R o n , s p of 34.36   m Ω c m 2 is acquired on a TMSBD with a W m e s a of 20 μm, inconsistent with the discussion above.
Figure 6b shows the forward J V characteristics of the regular SBD and 20–32 μm TMSBDs in semi-logarithmic scales. The subthreshold swing (SS) of 60.71 and 60.49 mV/dec and the ideality factors of 1.020 and 1.018 are extracted at TMSBDs and the regular SBD, respectively. Compared with the previous work, ideal factors of the TMSBDs in this research are competitive [19,29,30]. The ideality factors of both types of devices are close to unity, revealing that the carrier transport mechanism is dominated by thermionic emission (TE). The reverse saturation current density formula of the TE mechanism is J S = A * T 2 e q ϕ B / k T , where A * = 41.1   A c m 2 K 2 is the effective Richardson’s constant obtained when electron effective mass is m n * = 0.342 m 0 , T is Kelvin temperature, ϕ B is the Schottky barrier height, and k is the Boltzmann’s constant. The reverse saturation current densities of TMSBDs and the regular SBD are 5.92 × 10 8 and 8.76 × 10 8   A / c m 2 , respectively, and the extracted Schottky barrier heights are 0.82 and 0.81 eV, respectively. The barrier height of the TMSBDs is not much different from that of the regular SBD, which could be owing to the large W m e s a of the fabricated TMSBDs weakening the additional Schottky barrier introduced by the MOS junction of the sidewalls of the mesa. In addition, the barrier height of regular the SBD measured by J V measurement is 0.29 eV lower than that of the C V measurement, showing consistency with previously reported studies [1,27,36].
Figure 7 shows the I V characteristics of the 28 μm TMSBD and regular SBD under hysteresis measurements, which were obtained under the forward voltage sweeping from 0 to 5 V and then backward with a limiting maximum current of 100 mA. During the forward sweeping, the interface traps will capture electrons, which decreases the current growth rate and increases the turn-on voltage for the I V curve. However, during the backward sweeping, most of the electrons in the traps are too late to be released to capture new electrons, resulting in better I V characteristics of the device. Therefore, there is a hysteresis voltage ( V h ) in the forward and backward I V curves [38]. Additionally, the V h will increase with the interface trap density. For the regular SBD, I V hysteresis curves overlap completely, indicating that there are few interface traps in the device and the I V characteristics tend to be ideal. For the TMSBDs, the I V hysteresis curves show a higher V h , indicating that TMSBDs have a higher interface trap density compared with the regular SBD. It is seen that the V h here is only 40 mV, suggesting that the interface trap density of TMSBDs is very small. Due to the small interface trap density of the prepared TMSBDs, the ideality factor of the device is close to unity. In addition, the Schottky barrier height of β-Ga2O3 is primarily determined by the interface states rather than the actual metal [39]. Shallower trap levels increase the forward bias diode current and reduce the diode Schottky barrier height and turn-on voltage [40].
Figure 8a shows the statistics of extracted R o n , s p for the TMSBDs. The R o n , s p and A R of trench devices can be modeled as follows [27]:
R o n , s p = R w + R w / o R w A R
where R w is the R o n , s p below the mesa region of TMSBDs, and R w / o is the R o n , s p of the regular SBD with the same electrode shape and area as TMSBDs. R w and R w / o were determined to be 13.5077 and 16.77 m Ω c m 2 , respectively, from a fitting plot to the model (3) shown by the black dash line in Figure 8a. Additionally, the fitting degree of the plot R 2 is 0.9952, which is very close to 1, illustrating that the fitting result is reliable.
In practice, the electrons flow mainly through the mesa region, as shown in dashed blue lines in Figure 3a. The actual Schottky contact area is used to evaluate the current density as J o n = I o n / ( m · A s c h ) , where I o n is the forward current, A s c h is the Schottky contact area of one mesa, and m is the number of mesas in a device. The specific on-resistance R o n , s p based on the actual Schottky contact area was calculated using R o n , s p = d J o n / d V f 1 , where V f is the applied voltage. The extracted R o n , s p is statistically shown in Figure 8b. It has been predicted that the R o n , s p increases with increasing A R , which results from the current spreading mechanism. As the A R decreases, the contact area decreases, the current spreading weakens, and the current density increases, then the R o n , s p decreases [28].
Figure 9a depicts the representative reverse J V characteristics of the regular SBD and TMSBDs in semi-logarithmic scales. Owing to the small Schottky barrier height and the unrestricted barrier height lowering effect, the regular SBD has a large reverse leakage current and a small BV. Compared with the regular SBD, the TMSBDs had lower reverse leakage currents. This illustrates that the introduction of the trench MOS capacitance weakens the influence of the barrier height lowering effect, reducing the reverse leakage current of the devices. On the other hand, it suppresses the electric field crowding and improves the BV of the devices. The maximum BV of the TMSBDs was 340 V, which is 120 V higher than that of the regular SBD. Figure 9b shows the BV statistics of the TMSBDs. Unlike small-scale trench SBDs in the past research [27], the correlation between the BV and W m e s a of the large-scale TMSBDs has not been found. The electric field concentration is most prominent at the anode electrode edge [19]. It should be noted that introducing a thicker field-plate dielectric at the anode electrode edge can further enhance the BV.

4. Conclusions

In summary, we have designed and fabricated large-scale TMSBDs based on ultraviolet lithography and dry etching. The introduction of trench structures causes a deflection in the 1 / C 2 V plots. At the deflection point, the theoretically calculated depletion layer depth is 100 nm smaller than the actual trench depth, indicating that the sidewalls of the trench play an assisted role in depletion. It has been found that the hysteretic voltage between the forward and backward I V curves is small. This reveals that TMSBDs have a low interface trap density, contributing to a lower ideality factor of 1.02 and a turn-on voltage of 0.72 V. The specific on-resistance calculated with the anode area is inversely proportional to the A R , which is attributed to the increase in the current conduction path. The specific on-resistance determined by the actual Schottky contact area rises proportionally with the A R due to the current spreading phenomenon. In addition, compared with the regular SBD, the reverse performance of large-scale TMSBDs was greatly improved, which shows the important role of the trench MOS structure in suppressing the electric field crowding and barrier height lowering effect of β-Ga2O3 SBDs.

Author Contributions

Conceptualization, H.H. and W.L. (Wenjun Liu); methodology, H.H. and W.L. (Wenjun Liu); software, H.H.; validation, H.H., X.Z., Y.L. and W.L. (Wenjing Liu); formal analysis, H.H. and X.Z.; investigation, H.H., Y.L., W.L. (Wenjing Liu), H.Z. and G.X.; resources, H.H. and W.L. (Wenjun Liu); data curation, H.H, X.Z., Y.L., W.L. (Wenjing Liu), J.Y., H.Z. and G.X.; writing—original draft preparation, H.H.; writing—review and editing, H.H. and W.L. (Wenjun Liu); visualization, H.H., X.Z., Y.L., J.Y. and G.X.; supervision, X.Z. and Y.L.; project administration, H.H. and W.L. (Wenjun Liu); funding acquisition, W.L. (Wenjun Liu). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key R&D Program of China, grant number 2021YFB3202500.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, Z.; Gong, H.; Meng, C.; Yu, X.; Sun, X.; Zhang, C.; Ji, X.; Ren, F.; Gu, S.; Zheng, Y.; et al. Majority and Minority Carrier Traps in NiO/β-Ga2O3 p+-n Heterojunction Diode. IEEE Trans. Electron Devices 2022, 69, 981–987. [Google Scholar] [CrossRef]
  2. Orita, M.; Ohta, H.; Hirano, M.; Hosono, H. Deep-ultraviolet transparent conductive β-Ga2O3 thin films. Appl. Phys. Lett. 2000, 77, 4166–4168. [Google Scholar] [CrossRef]
  3. Higashiwaki, M.; Sasaki, K.; Kuramata, A.; Masui, T.; Yamakoshi, S. Gallium oxide (Ga2O3) metal-semiconductor field-effect transistors on single-crystal β-Ga2O3 (010) substrates. Appl. Phys. Lett. 2012, 100, 013504. [Google Scholar] [CrossRef]
  4. Mastro, M.A.; Kuramata, A.; Calkins, J.; Kim, J.; Ren, F.; Peartong, S.J. Opportunities and Future Directions for Ga2O3. ECS J. Solid State Sci. Technol. 2017, 6, P356–P359. [Google Scholar] [CrossRef]
  5. Villora, E.G.; Morioka, Y.; Atou, T.; Sugawara, T.; Kikuchi, A.; Fukuda, T. Infrared reflectance and electrical conductivity of β-Ga2O3. Phys. Status Solidi A-Appl. Res. 2002, 193, 187–195. [Google Scholar] [CrossRef]
  6. Kuramata, A.; Koshi, K.; Watanabe, S.; Yamaoka, Y.; Masui, T.; Yamakoshi, S. High-quality β-Ga2O3 single crystals grown by edge-defined film-fed growth. Jpn. J. Appl. Phys. 2016, 55, 1202A2. [Google Scholar] [CrossRef]
  7. Tomm, Y.; Reiche, P.; Klimm, D.; Fukuda, T. Czochralski grown Ga2O3 crystals. J. Cryst. Growth 2000, 220, 510–514. [Google Scholar] [CrossRef]
  8. Sasaki, K.; Higashiwaki, M.; Kuramata, A.; Masui, T.; Yamakoshi, S. Growth temperature dependences of structural and electrical properties of Ga2O3 epitaxial films grown on β-Ga2O3 (010) substrates by molecular beam epitaxy. J. Cryst. Growth 2014, 392, 30–33. [Google Scholar] [CrossRef]
  9. Wagner, G.; Baldini, M.; Gogova, D.; Schmidbauer, M.; Schewski, R.; Albrecht, M.; Galazka, Z.; Klimm, D.; Fornari, R. Homoepitaxial growth of β-Ga2O3 layers by metal-organic vapor phase epitaxy. Phys. Status Solidi A-Appl. Mat. 2014, 211, 27–33. [Google Scholar] [CrossRef]
  10. Baldini, M.; Albrecht, M.; Fiedler, A.; Irmscher, K.; Klimm, D.; Schewski, R.; Wagner, G. Semiconducting Sn-doped β-Ga2O3 homoepitaxial layers grown by metal organic vapour-phase epitaxy. J. Mater. Sci. 2016, 51, 3650–3656. [Google Scholar] [CrossRef]
  11. Konishi, K.; Goto, K.; Togashi, R.; Murakami, H.; Higashiwaki, M.; Kuramata, A.; Yamakoshi, S.; Monemar, B.; Kumagai, Y. Comparison of O2 and H2O as oxygen source for homoepitaxial growth of β-Ga2O3 layers by halide vapor phase epitaxy. J. Cryst. Growth 2018, 492, 39–44. [Google Scholar] [CrossRef]
  12. Togashi, R.; Nomura, K.; Eguchi, C.; Fukizawa, T.; Goto, K.; Quang Tu, T.; Murakami, H.; Kumagai, Y.; Kuramata, A.; Yamakoshi, S.; et al. Thermal stability of β-Ga2O3 in mixed flows of H2 and N2. Jpn. J. Appl. Phys. 2015, 54, 041102. [Google Scholar] [CrossRef]
  13. Gucmann, F.; Nadazdy, P.; Husekova, K.; Dobrocka, E.; Priesol, J.; Egyenes, F.; Satka, A.; Rosova, A.; Tapajna, M. Thermal stability of rhombohedral α- and monoclinic β-Ga2O3 grown on sapphire by liquid-injection MOCVD. Mater. Sci. Semicond. Process. 2023, 156, 107289. [Google Scholar] [CrossRef]
  14. Lv, Y.; Zhou, X.; Long, S.; Song, X.; Wang, Y.; Liang, S.; He, Z.; Han, T.; Tan, X.; Feng, Z.; et al. Source-Field-Plated β-Ga2O3 MOSFET With Record Power Figure of Merit of 50.4 MW/cm2. IEEE Electron Device Lett. 2019, 40, 83–86. [Google Scholar] [CrossRef]
  15. Lv, Y.; Liu, H.; Zhou, X.; Wang, Y.; Song, X.; Cai, Y.; Yan, Q.; Wang, C.; Liang, S.; Zhang, J.; et al. Lateral β-Ga2O3 MOSFETs With High Power Figure of Merit of 277 MW/cm2. IEEE Electron Device Lett. 2020, 41, 537–540. [Google Scholar] [CrossRef]
  16. Montgomery, R.H.; Zhang, Y.; Yuan, C.; Kim, S.; Shi, J.; Itoh, T.; Mauze, A.; Kumar, S.; Speck, J.; Graham, S. Thermal management strategies for gallium oxide vertical trench-fin MOSFETs. J. Appl. Phys. 2021, 129, 085301. [Google Scholar] [CrossRef]
  17. Zhou, X.; Ma, Y.; Xu, G.; Liu, Q.; Liu, J.; He, Q.; Zhao, X.; Long, S. Enhancement-mode β-Ga2O3 U-shaped gate trench vertical MOSFET realized by oxygen annealing. Appl. Phys. Lett. 2022, 121, 223501. [Google Scholar] [CrossRef]
  18. Konishi, K.; Goto, K.; Murakami, H.; Kumagai, Y.; Kuramata, A.; Yamakoshi, S.; Higashiwaki, M. 1-kV vertical Ga2O3 field-plated Schottky barrier diodes. Appl. Phys. Lett. 2017, 110, 103506. [Google Scholar] [CrossRef]
  19. Sasaki, K.; Wakimoto, D.; Thieu, Q.T.; Koishikawa, Y.; Kuramata, A.; Higashiwaki, M.; Yamakoshi, S. First Demonstration of Ga2O3 Trench MOS-Type Schottky Barrier Diodes. IEEE Electron Device Lett. 2017, 38, 783–785. [Google Scholar] [CrossRef]
  20. Hu, Z.; Zhou, H.; Kang, X.; Zhang, J.; Hao, Y.; Lv, Y.; Zhao, C.; Feng, Q.; Feng, Z.; Dang, K.; et al. Beveled Fluoride Plasma Treatment for Vertical β-Ga2O3 Schottky Barrier Diode With High Reverse Blocking Voltage and Low Turn-On Voltage. IEEE Electron Device Lett. 2020, 41, 441–444. [Google Scholar] [CrossRef]
  21. Roy, S.; Bhattacharyya, A.; Peterson, C.; Krishnamoorthy, S. 2.1 kV (001)-β-Ga2O3 vertical Schottky barrier diode with high-k oxide field plate. Appl. Phys. Lett. 2023, 122, 152101. [Google Scholar] [CrossRef]
  22. Neamen, D.A. Semiconductor Physics and Devices: Basic Principles, Fourth Edition, 4th ed.; Raghu Srinivasan: New York, NY, USA, 2012; pp. 338–339. [Google Scholar]
  23. Zhou, H.; Yan, Q.; Zhang, J.; Lv, Y.; Liu, Z.; Zhang, Y.; Dang, K.; Dong, P.; Feng, Z.; Feng, Q.; et al. High-Performance Vertical β-Ga2O3 Schottky Barrier Diode With Implanted Edge Termination. IEEE Electron Device Lett. 2019, 40, 1788–1791. [Google Scholar] [CrossRef]
  24. Wang, Y.; Cai, S.; Liu, M.; Lv, Y.; Long, S.; Zhou, X.; Song, X.; Liang, S.; Han, T.; Tan, X.; et al. High-Voltage (-201) β-Ga2O3 Vertical Schottky Barrier Diode With Thermally-Oxidized Termination. IEEE Electron Device Lett. 2020, 41, 131–134. [Google Scholar] [CrossRef]
  25. Yan, Q.; Gong, H.; Zhang, J.; Ye, J.; Zhou, H.; Liu, Z.; Xu, S.; Wang, C.; Hu, Z.; Feng, Q.; et al. β-Ga2O3 hetero-junction barrier Schottky diode with reverse leakage current modulation and BV2/Ron, sp value of 0.93 GW/cm2. Appl. Phys. Lett. 2021, 118, 122102. [Google Scholar] [CrossRef]
  26. Gong, H.H.; Yu, X.X.; Xu, Y.; Chen, X.H.; Kuang, Y.; Lv, Y.J.; Yang, Y.; Ren, F.F.; Feng, Z.H.; Gu, S.L.; et al. β-Ga2O3 vertical heterojunction barrier Schottky diodes terminated with p-NiO field limiting rings. Appl. Phys. Lett. 2021, 118, 202102. [Google Scholar] [CrossRef]
  27. Li, W.; Hu, Z.; Nomoto, K.; Jinno, R.; Zhang, Z.; Tu, T.Q.; Sasaki, K.; Kuramata, A.; Jena, D.; Xing, H.G. 2.44 kV Ga2O3 vertical trench Schottky barrier diodes with very low reverse leakage current. In Proceedings of the IEEE Annual International Electron Devices Meeting (IEDM), San Francisco, CA, USA, 1–5 December 2018. [Google Scholar] [CrossRef]
  28. Li, W.; Nomoto, K.; Hu, Z.; Jen, D.; Xing, H.G. Fin-channel orientation dependence of forward conduction in kV-class Ga2O3 trench Schottky barrier diodes. Appl. Phys. Express 2019, 12, 061007. [Google Scholar] [CrossRef]
  29. Li, W.; Nomoto, K.; Hu, Z.; Jena, D.; Xing, H.G. Field-Plated Ga2O3 Trench Schottky Barrier Diodes With a BV2/Ron, sp of up to 0.95 GW/cm2. IEEE Electron Device Lett. 2020, 41, 107–110. [Google Scholar] [CrossRef]
  30. Jian, Z.; Mohanty, S.; Ahmadi, E. Temperature-dependent current-voltage characteristics of β-Ga2O3 trench Schottky barrier diodes. Appl. Phys. Lett. 2020, 116, 152104. [Google Scholar] [CrossRef]
  31. Wilhelmi, F.; Kunori, S.; Sasaki, K.; Kuramata, A.; Komatsu, Y.; Lindemann, A. Packaged β-Ga2O3 Trench MOS Schottky Diode With Nearly Ideal Junction Properties. IEEE Trans. Power Electron. 2022, 37, 3737–3742. [Google Scholar] [CrossRef]
  32. Mehrotra, M.; Baliga, B.J. The trench MOS barrier Schottky (TMBS) rectifier. In Proceedings of the IEEE International Electron Devices Meeting, Washington, DC, USA, 5–8 December 1993. [Google Scholar] [CrossRef]
  33. Zhang, Y.; Sun, M.; Liu, Z.; Piedra, D.; Hu, J.; Gao, X.; Palacios, T. Trench formation and corner rounding in vertical GaN power devices. Appl. Phys. Lett. 2017, 110, 193506. [Google Scholar] [CrossRef]
  34. Yuekseltuerk, E.; Bengi, S. The frequency dependent complex dielectric and electric modulus properties of Au/P3HT/n-Si (MPS) Schottky barrier diode (SBD). J. Mater. Sci. Mater. Electron. 2023, 34, 1–13. [Google Scholar] [CrossRef]
  35. Dong, H.; Mu, W.; Hu, Y.; He, Q.; Fu, B.; Xue, H.; Qin, Y.; Jian, G.; Zhang, Y.; Long, S.; et al. C-V and J-V investigation of HfO2/Al2O3 bilayer dielectrics MOSCAPs on (OW) β-Ga2O3. AIP Adv. 2018, 8, 065215. [Google Scholar] [CrossRef]
  36. Sasaki, K.; Higashiwaki, M.; Kuramata, A.; Masui, T.; Yamakoshi, S. Ga2O3 Schottky Barrier Diodes Fabricated by Using Single-Crystal β-Ga2O3 (010) Substrates. IEEE Electron Device Lett. 2013, 34, 493–495. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Zhang, J.; Zhou, H.; Zhang, T.; Wang, H.; Feng, Z.; Hao, Y. Leakage current mechanisms of groove-type tungsten-anode GaN SBDs with ultra low turn-on voltage and low reverse current. Solid-State Electron. 2020, 169, 107807. [Google Scholar] [CrossRef]
  38. Hao, W.; He, Q.; Zhou, K.; Xu, G.; Xiong, W.; Zhou, X.; Jian, G.; Chen, C.; Zhao, X.; Long, S. Low defect density and small I–V curve hysteresis in NiO/β-Ga2O3 pn diode with a high PFOM of 0.65 GW/cm2. Appl. Phys. Lett. 2021, 118, 043501. [Google Scholar] [CrossRef]
  39. Fu, B.; Jian, G.; He, G.; Feng, B.; Mu, W.; Li, Y.; Jia, Z.; Li, Y.; Long, S.; Ding, S.; et al. Investigation on β-Ga2O3 (101) plane with high-density surface dangling bonds. J. Alloys Compd. 2021, 889, 161714. [Google Scholar] [CrossRef]
  40. Surdi, H.; Koeck, F.A.M.; Ahmad, M.F.; Thornton, T.J.; Nemanich, R.J.; Goodnick, S.M. Demonstration and Analysis of Ultrahigh Forward Current Density Diamond Diodes. IEEE Trans. Electron Devices 2022, 69, 254–261. [Google Scholar] [CrossRef]
Figure 1. Cross-sectional schematic of the β-Ga2O3 TMSBDs.
Figure 1. Cross-sectional schematic of the β-Ga2O3 TMSBDs.
Electronics 12 04315 g001
Figure 2. Process flow of the fabricated TMSBDs.
Figure 2. Process flow of the fabricated TMSBDs.
Electronics 12 04315 g002
Figure 3. Surface optical micrograph of (a) the 24 μm β-Ga2O3 TMSBD and (b) regular SBD.
Figure 3. Surface optical micrograph of (a) the 24 μm β-Ga2O3 TMSBD and (b) regular SBD.
Electronics 12 04315 g003
Figure 4. (a) C V characteristics and (b) 1 / C 2 V plots of the regular SBD and TMSBDs.
Figure 4. (a) C V characteristics and (b) 1 / C 2 V plots of the regular SBD and TMSBDs.
Electronics 12 04315 g004
Figure 5. Frequency dependences (a) C V characteristics and (b) 1 / C 2 V plots of the 28 μm TMSBD.
Figure 5. Frequency dependences (a) C V characteristics and (b) 1 / C 2 V plots of the 28 μm TMSBD.
Electronics 12 04315 g005
Figure 6. Forward J V characteristics of the regular SBD and TMSBDs in (a) linear scales and (b) semi-logarithmic scales.
Figure 6. Forward J V characteristics of the regular SBD and TMSBDs in (a) linear scales and (b) semi-logarithmic scales.
Electronics 12 04315 g006
Figure 7. The I V characteristics of the 28 μm TMSBD and regular SBD under hysteresis measurements.
Figure 7. The I V characteristics of the 28 μm TMSBD and regular SBD under hysteresis measurements.
Electronics 12 04315 g007
Figure 8. Statistics of the extracted (a) R o n , s p and (b) R o n , s p for the TMSBDs.
Figure 8. Statistics of the extracted (a) R o n , s p and (b) R o n , s p for the TMSBDs.
Electronics 12 04315 g008
Figure 9. (a) Representative reverse J V characteristics of the regular SBD and TMSBDs in semi-logarithmic scales; (b) BV statistics of the TMSBDs.
Figure 9. (a) Representative reverse J V characteristics of the regular SBD and TMSBDs in semi-logarithmic scales; (b) BV statistics of the TMSBDs.
Electronics 12 04315 g009
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

He, H.; Zhou, X.; Liu, Y.; Liu, W.; Yang, J.; Zhang, H.; Xie, G.; Liu, W. Large-Scale β-Ga2O3 Trench MOS-Type Schottky Barrier Diodes with 1.02 Ideality Factor and 0.72 V Turn-On Voltage. Electronics 2023, 12, 4315. https://doi.org/10.3390/electronics12204315

AMA Style

He H, Zhou X, Liu Y, Liu W, Yang J, Zhang H, Xie G, Liu W. Large-Scale β-Ga2O3 Trench MOS-Type Schottky Barrier Diodes with 1.02 Ideality Factor and 0.72 V Turn-On Voltage. Electronics. 2023; 12(20):4315. https://doi.org/10.3390/electronics12204315

Chicago/Turabian Style

He, Hao, Xinlong Zhou, Yinchi Liu, Wenjing Liu, Jining Yang, Hao Zhang, Genran Xie, and Wenjun Liu. 2023. "Large-Scale β-Ga2O3 Trench MOS-Type Schottky Barrier Diodes with 1.02 Ideality Factor and 0.72 V Turn-On Voltage" Electronics 12, no. 20: 4315. https://doi.org/10.3390/electronics12204315

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