Next Article in Journal
Band Broadening of Terahertz Photonic Crystals Circulator Using Two Symmetrical Hexagonal Aluminum Sheets
Previous Article in Journal
Phase-Field Crystal Studies on Grain Boundary Migration, Dislocation Behaviors, and Topological Transition under Tension of Square Polycrystals
Previous Article in Special Issue
Preparation of Ti-Doped ZnO/Bi2O3 Nanofilm Heterojunction and Analysis of Microstructure and Photoelectric Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Low On–State Resistance Si/4H–SiC Heterojunction VDMOS with Electron Tunneling Layer Based on a Discussion of the Hetero–Transfer Mechanism

1
The State Key Laboratory of Electronic Thin Films and Integrated Devices, School of Integrated Circuit Science and Engineering, University of Electronic Science and Technology of China, Chengdu 610000, China
2
China Zhenhua Group Yongguang Electronics Co., Ltd., Guiyang 550000, China
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(5), 778; https://doi.org/10.3390/cryst13050778
Submission received: 31 March 2023 / Revised: 4 May 2023 / Accepted: 5 May 2023 / Published: 7 May 2023
(This article belongs to the Special Issue Nano-Semiconductors: Devices and Technology)

Abstract

:
In this study, we propose a novel silicon (Si)/silicon carbide (4H–SiC) heterojunction vertical double–diffused MOSFET with an electron tunneling layer (ETL) (HT–VDMOS), which improves the specific on–state resistance (RON), and examine the hetero–transfer mechanism by simulation. In this structure, the high channel mobility and high breakdown voltage (BV) are obtained simultaneously with the Si channel and the SiC drift region. The heavy doping ETL on the 4H–SiC side of the heterointerface leads to a low heterointerface resistance (RH), while the RH in H–VDMOS is extremely high due to the high heterointerface barrier. The higher carrier concentration of the 4H–SiC surface can significantly reduce the width of the heterointerface barrier, which is demonstrated by the comparison of the conductor energy bands of the proposed HT–VDMOS and the general Si/SiC heterojunction VDMOS (H–VDMOS), and the electron tunneling effect is significantly enhanced, leading to a higher tunneling current. As a result, a significantly improved trade–off between RON and BV is achieved. With similar BV values (approximately 1525 V), the RON of the HT–VDMOS is 88% and 65.75% lower than that of H–VDMOS and the conventional SiC VDMOS, respectively.

1. Introduction

Silicon carbide (4H–SiC) MOSFET shows excellent performance in the applications of high power, high frequency and high temperature owing to the prominent material characteristics of 4H–SiC [1]. As a unipolar device, 4H–SiC MOSFET possesses a faster switching speed than the Si IGBT while achieving a much higher breakdown voltage compared with the Si MOSFET [2,3]. The critical electric field of SiC is almost 10 times that of Si, resulting in a much thinner thickness of the drift region of the SiC MOSFET with the same withstanding voltage, namely, a much lower drift resistance. However, the low channel mobility caused by the high SiC/SiO2 interface state density which is introduced in the gate oxide fabricated processes is the most important limit for the development of SiC MOSFET [4]. Channel resistance is the major component of the specific on–state resistance (RON) of SiC MOSFET. Moreover, the poor–quality gate oxide of SiC MOSFET leads to some stability issues, such as threshold voltage (VTH) shift [5]. Therefore, research into the suppression of the SiC/SiO2 interface states and improving the gate oxide quality are pivotal to the SiC MOSFET.
In the past two decades, studies have been carried out on the SiC/SiO2 interface, and many process optimization method have been used to decrease the density of the interface state in SiC/SiO2 [6,7,8,9]. Nevertheless, the channel mobility of SiC MOSFET is still much lower than that of Si MOSFET (above 50% of the bulk mobility of 1360 cm2/Vs) [10]. Attempts to address the channel mobility issue in SiC MOSFET need further research and experiments. As the Si channel has higher mobility, the Si/SiC heterojunction vertical double–diffused MOSFET (VDMOS) device, which combines the high mobility of the Si channel and the high critical electronic field of the SiC drift region, was proposed by L. Chen et al. [11]. Baoxing Duan et al. proposed a Si/SiC heterojunction VDMOS with breakdown point transfer technology to improve the reverse stability of the device [12]. The Si/SiC heterojunction structure has great potential in vertical power devices, especially in power MOSFETs. However, the transfer mechanisms of the Si/SiC heterojunction are not clear, and the forward characteristics of the Si/SiC heterojunction VDMOS can be further optimized.
In this paper, a novel Si/4H–SiC heterojunction VDMOS with an electron tunnel layer (ETL) (HT–VDMOS) is proposed. On the 4H–SiC side of the heterointerface in the JFET region of the proposed structure, a thin heavy doping layer, which we call ETL, is introduced to increase the tunneling current of the heterojunction at on–state. The heterojunction transfer mechanism analysis indicates that the higher doping concentration of the 4H–SiC surface can dramatically decrease the heterointerface resistance (RH) due to the much narrower heterointerface barrier, namely, the greatly higher tunnel current. Consequently, a significantly lower RON is achieved by the HT–VDMOS compared with the general Si/4H–SiC heterojunction VDMOS (H–VDMOS) and the conventional SiC VDMOS (C–VDMOS). In addition, the results of the simulation indicate that the heterointerface states benefit the RON but increase the gate charge (QG) and gate–to–drain charge (QGD) of the HT–VDMOS.

2. Device Structure and Transfer Mechanism

2.1. HT–VDMOS Device Structure

The structures of the HT–VDMOS, H–VDMOS and C–VDMOS are shown in Figure 1. Both heterojunction devices take advantage of the high channel mobility of Si and the high withstanding voltage of 4H–SiC. The ETL on the SiC side of the heterojunction of the JFET region in the HT–VDMOS can significantly enhance the electron tunnel current of the heterojunction, and this results in a lower RON. As the ETL has a negative effect on the electric field in the gate oxide and the Si PN junction, the lower width of the JFET (WJFET) region is required for the HT–VDMOS. WJFET values of 2 μm, 3 μm and 3 μm were chosen for the HT–VDMOS, H–VDMOS and C–VDMOS, respectively, while the doping concentration, thickness and length of the ETL in HT–VDMOS were optimized to be 5 × 1017 cm−3, 0.07 μm and 1 μm, respectively. The thickness of the Si layer in HT–VDMOS and H–VDMOS was 0.1 μm. The doping concentration of the channels of the HT–VDMOS, H–VDMOS and C–VDMOS was set to 9.2 × 1016 cm−3, 9.2 × 1016 cm−3 and 1.7 × 1016 cm−3, respectively. The channel mobility of the C–VDMOS was set to 20 cm2/Vs.

2.2. Heterojunction Transfer Mechanism

Owing to the different band gap and electron affinity between Si and 4H–SiC, a band offset occurs in the heterointerface, as shown in Figure 2a, and a high interface barrier similar to the Schottky contact leads to a high RH when setting a positive bias on the 4H–SiC side (drain electrode). As a result, the RON of the heterojunction device without the ETL is significantly increased due to the large value of the RH. In an attempt to reduce the RH, the ETL structure was introduced into the novel HT–VDMOS. As the energy band of the heterojunction with the ETL on the SiC side shows in Figure 2b, a higher doping concentration decreases the energy level in the ETL and a lower width of the barrier in the heterointerface is obtained.
There are two transport modes for electrons in the heterojunction: the thermionic emission mode and the direct tunneling mode. In the thermionic emission mode shown in Figure 1a, the electrons require enough energy to cross the high heterointerface barrier, resulting in a low current and a high RH. When the electrons transfer by direct tunneling, the tunneling probability (PT) is determined by the tunneling distance (DT). Based on the Wentzel–Kramers–Brillouin approximation, the relation between PT and DT for the triangular barrier can be evaluated by the following formula:
P T exp ( 4 2 m * ψ b 3 h D T )
where m * is the effective mass, ψ b is the triangular barrier height and h is the Planck constant. PT is the exponential positive correlated with DT, and the direct tunneling becomes the main transfer mode when the width of the barrier reduces to a small value, as shown in Figure 2b. In this condition, the heterojunction exhibits ohmic behavior, and a large tunneling current can be obtained. Referring to the current of the ohmic contact, the direct tunneling current of the heterojunction (IHT) in the HT–VDMOS can be analyzed by the following formula:
I H T V D S exp { 4 π h m n * ε r ε 0 ( ψ S i C n E T L ) }
where m n * is the electron effective mass, ε r ε 0 is the permittivity of 4H–SiC, ψ S i C is the heterointerface barrier height and n E T L is the concentration of electrons in the ETL. Since the ψ S i C shows a weak correlation with the doping concentration of ETL (NETL) and IHT is strongly and positively correlated to the n E T L , a higher NETL can lead to a higher IHT. Therefore, the ETL structure can significantly decrease the RON of the HT–VDMOS due to the much lower RH.
The simulation results demonstrate the analysis of the electron transfer mechanism of the heterojunction. Figure 3a shows the simulation results for the conductor energy band (EC) with different NETL values. The EC drops on the 4H–SiC side of the heterointerface (in the ETL) with when NETL increases, leading to the decrease in the width of the heterointerface barrier. As the NETL increases to 5 × 1017 cm−3, a narrow heterointerface barrier, through which the electrons can easily tunnel, is obtained at the heterojunction in the HT–VDMOS, and both a high tunneling current and a low RH are achieved.
The influence of the thickness of the ETL (TETL) on the EC of the heterojunction according to the simulation is shown in Figure 3b. The TETL should not be excessively small, as with a much thinner ETL, the energy band on the SiC side cannot drop enough to form the narrow heterointerface barrier for the electron tunneling.

2.3. Heterointerface Charge Analysis

To date, two technologies, heteroepitaxy growth and wafer bonding, can be utilized to realize the Si/SiC heterojunction devices [13,14,15,16,17]. Both technologies will generate dislocations and defects at the heterointerface owing to the 19.3% lattice mismatch value between the Si and SiC. This heterointerface state can affect the heterojunction on the I–V and C–V characteristics [18,19]. The current of the heterojunction can be increased by the interface states, which we consider to be caused by the trap–assisted tunneling shown in Figure 4. In the barrier, electrons can tunnel into the defect energy level provided by the heterointerface states before their transfer to the SiC, and a lower electron DT and a higher tunneling current are obtained.
Some studies have utilized the donor–type heterointerface charges (electrons) to investigate the effect of heterointerface dislocations by simulation [20,21]. Similar to trap–assisted tunneling, the donor–type heterointerface charges can reduce the DT for electrons and lead to a significant increase in the forward current of the HT–VDMOS. The heterointerface charge defaults to the donor type in the rest of this paper.

3. Device Simulation and Discussion

3.1. Simulation Models and Mobility Parameters

The following physics models were utilized in the TCAD simulation: cvt, analytic, conmob, fldmob, srh, incomplete, bgn, auger, fermi, optr and hei. The cvt, analytic, conmob and fldmob models are correlated to the mobility parameters. The parameters for the mobility of Si are set to be the default of the Silvaco, and the core parameters of the channel–carrier mobility of 4H–SiC are listed in Table 1. We increased the surface roughness factor of the 4H–SiC and the channel–carrier mobility was reduced to 20 cm2/Vs. In contrast, the channel–carrier mobility of Si was 500 cm2/Vs, according to the simulation.

3.2. Optimization of Structure Characteristics

The ETL structure can significantly improve the forward performance of the HT–VDMOS, but this also results in reverse stability issues, such as the increasing electric field at the gate oxide and the PN junction in Si layer. In order to ensure the ability of the withstand voltage of the HT–VDMOS, a detailed study of the structure optimization was carried out via the TCAD Silvaco to achieve a low RON and high reverse reliability simultaneously for the proposed device.
We studied the influences of four important structure characteristics of the HT–VDMOS on the RON, the maximum electric field of the gate oxide (EOX–M) and the electric field at the Si PN junction (ESi). The influence of NETL and TETL on the RON, EOX–M and ESi is shown in Figure 5a,b. In agreement with the analysis of the EC, as the NETL and TETL values increase, the RON decreases while the EOX–M and ESi values increase. When the NETL and TETL decrease to small values, the RON of the HT–VDMOS increases rapidly due to the failure of the electron tunneling. The variation of the EOX–M is more obvious than that of the ESi under the effect of the NETL and TETL, and the ESi never exceeds the critical electric field of Si. Thus, considering the trade–off between the RON and EOX–M, the NETL value of 5 × 1017 cm−3 and the TETL value of 0.07 μm were chosen as the optimized values for the HT–VDMOS.
The length of the ETL (LETL) also remarkably affects the device characteristics, as shown in Figure 5c. The influence on the EOX–M is weak, but the larger LTEL will significantly increase the ESi. Furthermore, the improvement in the RON is less dramatic with larger values of the LETL. Therefore, the optimized value of the LTEL in the HT–VDMOS was chosen to be 1 μm.
The last structure characteristic we studied was the WJFET, and the simulation results are shown in Figure 5d. Since the introduction of the ETL leads to a higher doping concentration of the JFET region, a lower WJFET is required for the HT–VDMOS than the C–VDMOS. The WJFET shows the significant influence on the EOX–M, and the RON is also increased with the reduction of the WJFET owing to the larger resistance of the JFET region (RJ). Synthesizing the effects of the WJFET on the EOX–M and RON, we chose a final WJFET value of 2 μm for the HT–VDMOS.

3.3. Comparison of Device Performances

Among the HT–VDMOS, H–VDMOS and C–VDMOS devices, the proposed HT–VDMOS achieves the lowest RON due to the lower channel resistance and RH, as shown in Figure 6. The HT–VDMOS and H–VDMOS have a much lower RCH than the C–VDMOS, owing to the high channel mobility, while RH value of the C–VDMOS is zero. The RH, which is extremely high in the H–VDMOS, can be greatly reduced by the ETL structure in the proposed HT–VDMOS. In spite of the narrower JFET region of the HT–VDMOS leading to a higher RJFET, the increased values of the JFET resistance (RJFET) and RH are significantly lower than the decreased value of the RCH compared with the C–VDMOS, resulting in the lowest RON for the HT–VDMOS among these three structures. On the contrary, the H–VDMOS shows a poor forward current ability due to the extremely high RH.
The breakdown voltage of the HT–VDMOS is slightly higher than those of the H–VDMOS and C–VDMOS. As discussed before, the EOX–M and ESi are limited below the safety values, and the breakdown point of the HT–VDMOS is the P + N junction in the 4H–SiC region. Thus, the BV of the VDMOS devices is determined by the electric field of the edge of the P–well region. According to the curvature effect, the lower WJFET in the HT–VDMOS leads to a large curvature and results in the weakened concentration of the electric field, namely, a higher BV.
The HT–VDMOS possesses the lowest miller capacitance (Crss) among these three devices at VD = 400 V, while the Crss of the H–VDMOS is the highest, as shown in Figure 7. The Crss of the VDMOS device is determined by the barrier capacitance in the JFET and drift regions (CB) and the gate oxide capacitance (COX). At the condition of high VD voltage bias, the COX is the major component of the Crss, and the lower WJFET of the HT–VDMOS leads to a lower Crss than the other two devices. The permittivity of the Si layer in the HT–VDMOS and H–VDMOS is higher than the 4H–SiC, resulting in a higher CB in the JFET region than in the C–VDMOS. Therefore, the Crss of the H–VDMOS is highest at VD = 400 V.
At the condition of low VD voltage bias, the CB rapidly increases with the decrease of the VD and becomes the main part of the Crss. As a higher CB is obtained for the HT–VDMOS and H–VDMOS due to the higher permittivity of the Si layer, the Crss of the C–VDMOS is the lowest at this condition. The ETL also leads to a higher CB owing to the narrower width of the barrier region caused by the higher doping concentration. Consequently, an abnormal increase occurs in the Crss–VD curve of the HT–VDMOS with a VD value from 1 V to 9 V.
Figure 8 shows the gate charge tests of HT–VDMOS, H–VDMOS and C–VDMOS devices. The HT–VDMOS and C–VDMOS obtain a similar QG, which is lower than that of the H–VDMOS. The QGD of the HT–VDMOS is the lowest among the three devices, which is attributed to the lowest Crss at a high voltage bias condition. Relatively, the QGD of the H–VDMOS is higher than that of the HT–VDMOS and C–VDMOS. The higher gate platform voltage (VGP) of the C–VDMOS is caused by the higher VTH and the lower μ C H . The VGP can be expressed by the following formula:
V G P = V T H + J O N W C e l l L C H 2 μ C H C O X
where J O N is the on–state current density, W C e l l is the cell width, L C H is the length of the channel and μ C H is the channel mobility. The comparison of the main characteristics of the HT–VDMOS, H–VDMOS and C–VDMOS is listed in Table 2.

3.4. Analysis and Discussion of RON

The RON of the VDMOS device consists of five components: the source resistance RS, RCH, RH, the RJFET, and the drain resistance RND, as shown in Figure 9a. According to the analysis of the components shown in Figure 9b, the major differences in the total RON originate from the RCH, RH and RJFET, which has been discussed previously. With the increasing VG, the RCH of these two structures both significantly decrease due to the higher channel carrier density. However, the proportion of the RCH in the RON of the HT–VDMOS is low, and the variation of RCH caused by VG is not obvious for the RON. On the contrary, as the proportion of the RCH in the RON is much higher for the C–VDMOS due to the poor channel mobility, the reduction in RON caused by the increasing VG is much more significant, although the increase in the channel carrier density of the C–VDMOS with the increasing VG is slower than that of the HT–VDMOS owing to the higher width of band gap of SiC.
The influence of the charge density in the gate oxide (QOX) on the RON of these two structures is similar to that of the VG, as shown in Figure 10a. During the practical work of VDMOS, a few charges will enter the gate oxide, leading to a threshold voltage shift after long–term work. Thus, the different values of QOX can also affect the channel carrier density, resulting in a similar variation in the RON of these two structures compared with the VG.
The heterojunction can be fabricated by wafer bonding or deposition. If the Si deposition process was utilized, the Si film may not be monocrystalline, and the RON of the HT–VDMOS will be different when the Si film is polycrystalline due to the variation of the μ C H . Therefore, the influence of the μ C H of the HT–VDMOS on the RON was studied, as shown in Figure 10b. The RON of the HT–VDMOS decreases with degeneration of the μ C H and the decrease becomes rapid when the μ C H of HT–VDMOS is lower than 150 cm2/Vs. Nevertheless, the proposed HT–VDMOS can achieve a lower RON than the C–VDMOS, as long as the μ C H of the HT–VDMOS is higher than 69.8 cm2/Vs.

3.5. Influence of Heterointerface Charges

The influence of the heterointerface charge density (QI) on the output I–V characteristic and RON of the HT–VDMOS is shown in Figure 11. The forward current of the HT–VDMOS is enhanced with increasing QI, which is caused by the trap–assisted tunneling. The RON values of 3.67 mΩ × cm2, 3.64 mΩ × cm2, 3.61 mΩ × cm2, 3.59 mΩ × cm2 and 3.52 mΩ × cm2 were obtained for the HT–VDMOS for QI = 0, 1 × 1011 cm−3, 3 × 1011 cm−3, 1 × 1012 cm−3 and 3 × 1012 cm−3, respectively. The analysis of the components of the RON of HT–VDMOS with different QI values, shown in Figure 11b, demonstrated that trap–assisted tunneling can reduce the RH, resulting in a lower RON. As the proportion of the RH in RON is also low, the variation of RON is slight, and the reduction of the RON caused by the heterointerface charges becomes slow when the QI values become larger, as shown in the inset graph in Figure 11a. These results indicate that the RH of the HT–VDMOS is becomes lower as a minor component of a RON with a high QI.
The QG and QGD of the HT–VDMOS increase with increasing QI, as shown in Figure 12. The heterointerface charges lead to a higher Crss due to the narrower width of the barrier in the JFET region, and this causes extra charges for the gate charges, resulting in a higher QGD and QG in the HT–VDMOS. QGD values of 54 nC, 57 nC, 60 nC, 69 nC and 85 nC and QG values of 371 nC, 369 nC, 368 nC, 379 nC and 401 nC were obtained for the HT–VDMOS for QI = 0, 1 × 1011 cm−3, 3 × 1011 cm−3, 1 × 1012 cm−3 and 3 × 1012 cm−3, respectively.

4. Conclusions

A novel HT–VDMOS structure, which features an introduced ETL and a low RON, is proposed, and we discuss its hetero–transfer mechanism based on device simulation studies. The HT–VDMOS possesses the advantages of Si and 4H–SiC, namely, high channel mobility and high critical electric field, respectively. The heavily doped ETL on the 4H–SiC side of the heterointerface in the JFET region can greatly reduce the RH caused by the high barrier in the heterointerface due to the direct tunneling effect. With a high doping concentration, the energy band level on the 4H–SiC side of the heterointerface can be decreased, and a significantly narrower interface barrier is formed. Thus, the electrons on the Si side can easily transfer to 4H–SiC by direct tunneling, and a lower RH is achieved. Our study of the influence of the heterointerface charges indicates that the heterointerface charges can enhance the forward current of the HT–VDMOS but leads to a degeneration of gate charge characteristics. Consequently, the RON of the HT–VDMOS is decreased by 88% and 65.8% compared with the H–VDMOS and C–VDMOS, respectively.

Author Contributions

Conceptualization, H.C.; methodology, H.C.; validation, C.L.; formal analysis, Y.Z.; investigation, H.C; resources, Z.W.; writing—original draft preparation, H.C.; writing—review and editing, Y.Z.; supervision, B.Z. and Z.L.; project administration, Y.Z.; funding acquisition, R.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Major Science and Technology Program of Anhui Province under Grant No. 2020b05050007.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The study did not report any data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Baliga, B.J. Silicon Carbide Power Devices; World Scientific: Singapore, 2005. [Google Scholar]
  2. Minamisawa, R.A.; Vemulapati, U.; Mihaila, A.; Papadopoulos, C.; Rahimo, M. Current Sharing Behavior in Si IGBT and SiC MOSFET Cross–Switch Hybrid. IEEE Electron Device Lett. 2016, 37, 1178–1780. [Google Scholar] [CrossRef]
  3. Konstantinov, A.; Wahab, Q.; Nordell, N.; Lindefelt, U. Ionization rates and critical fields in 4H silicon carbide. Appl Phys Lett. 1997, 71, 90–92. [Google Scholar] [CrossRef]
  4. Kimoto, T.; Cooper, J.A. Fundamentals of Silicon Carbide Technology: Growth, Characterization, Devices, and Applications; Wiley: Singapore, 2014; pp. 212–248. [Google Scholar]
  5. Yano, H.; Kanafuji, N.; Osawa, A.; Hatayama, T.; Fuyuki, T. Threshold Voltage Instability in 4H–SiC MOSFETs with Phosphorus–Doped and Nitrided Gate Oxides. IEEE Trans. Electron Devices 2015, 62, 324–332. [Google Scholar] [CrossRef]
  6. Chung, G.Y.; Tin, C.C.; Williams, J.R.; McDonald, K.; Chanana, R.K.; Weller, R.A.; Pantelides, S.T.; Feldman, L.C.; Holland, O.W.; Das, M.K.; et al. Improved inversion channel mobility for 4H–SiC MOSFETs following high temperature anneals in nitric oxide. IEEE Electron Device Lett. 2001, 22, 176–178. [Google Scholar] [CrossRef]
  7. Poggi, A.; Moscatelli, F.; Solmi, S.; Nipoti, R. Investigation on the Use of Nitrogen Implantation to Improve the Performance of N–Channel Enhancement 4H–SiC MOSFETs. IEEE Trans. Electron Devices 2008, 55, 2021–2028. [Google Scholar] [CrossRef]
  8. Nanen, Y.; Kato, M.; Suda, J.; Kimoto, T. Effects of Nitridation on 4H–SiC MOSFETs Fabricated on Various Crystal Faces. IEEE Trans. Electron Devices 2013, 60, 1260–1262. [Google Scholar] [CrossRef]
  9. Yang, X.; Lee, B.; Misra, V. Electrical Characteristics of SiO2 Deposited by Atomic Layer Deposition on 4H–SiC After Nitrous Oxide Anneal. IEEE Trans. Electron Devices 2016, 63, 2826–2830. [Google Scholar] [CrossRef]
  10. Noguchi, M.; Iwamatsu, T.; Amishiro, H.; Watanabe, H.; Kita, K.; Yamakawa, S. Determination of intrinsic phonon–limited mobility and carrier transport property extraction of 4H–SiC MOSFETs. In Proceedings of the 2017 IEEE International Electron Devices Meeting (IEDM), San Francisco, CA, USA, 2–6 December 2017. [Google Scholar]
  11. Chen, L.; Guy, O.J.; Jennings, M.R.; Igic, P.; Wilks, S.P.; Mawby, P.A. Study of a novel Si/SiC hetero–junction MOSFET. Solid–State Electron. 2007, 51, 662–666. [Google Scholar] [CrossRef]
  12. Duan, B.; Lv, J.; Zhao, Y.; Yang, Y. SiC/Si heterojunction VDMOS breaking silicon limit by breakdown point transfer technology. Micro Nano Lett. 2016, 13, 96–99. [Google Scholar] [CrossRef]
  13. Ikoma, Y.; Endo, T.; Watanabe, F.; Motooka, T. Growth of Si/3C–SiC/Si (100) heterostructures by pulsed supersonic free jets. Appl. Phys. Lett. 1999, 75, 3977–3979. [Google Scholar] [CrossRef]
  14. Zheng, H.; Zhu, J.; Fu, Z.; Lin, B.; Li, X. Heteroepitaxial growth and characterization of 3C–SiC films on Si substrates using LPVCVD. J. Mater. Sci. Technol. 2005, 21, 536–540. [Google Scholar]
  15. Hens, P.; Brow, R.; Robinson, H.; Cromar, M.; Zeghbroeck, B.V. Epitaxial growth of cubic silicon carbide on silicon using hot filament chemical vapor deposition. Thin Solid Films 2017, 635, 48–52. [Google Scholar] [CrossRef]
  16. Xu, Q.; Zhu, P.; Sun, Q.; Tu, R.; Yang, M.; Zhang, S.; Zhang, L.; Goto, T.; Yan, J.; Li, S. Elimination of double position domains (DPDs) in epitaxial < 111 >–3C–SiC on Si (111) by laser CVD. Appl. Surf. Sci. 2017, 426, 662–666. [Google Scholar]
  17. Nishida, S.; Liang, J.; Hayashi, T.; Morimoto, M.; Shigekawa, N.; Arai, M. Annealing characteristics of p+–Si/n–4H–SiC junctions by using Surface–Activated bonding. In Proceedings of the 2014 4th IEEE International Workshop on Low Temperature Bonding for 3D Integration (LTB–3D), Tokyo, Japan, 15–16 July 2014. [Google Scholar]
  18. Liang, J.; Nishida, S.; Hayashi, T.; Arai, M.; Shigekawa, N. Effects of interface state charges on the electrical properties of Si/SiC heterojunctions. Appl. Phys. Lett. 2014, 105, 151607. [Google Scholar] [CrossRef]
  19. Liang, J.; Nishida, S.; Arai, M.; Shigekawa, N. Improved electrical properties of n–n and p–n Si/SiC junctions with thermal annealing treatment. J. Appl. Phys. 2016, 120, 034504. [Google Scholar] [CrossRef]
  20. Wang, Y.; Wang, H.; Cao, F.; Wang, H. High Performance of Polysilicon/4H–SiC Dual–Heterojunction Trench Diode. IEEE Trans. Electron Devices 2017, 64, 1653–1659. [Google Scholar] [CrossRef]
  21. Sun, L.; Duan, B.; Yang, X.; Wang, Y.; Yang, Y. Analysis of the novel Si/SiC heterojunction IGBT characteristics by TCAD simulation. Superlattices Microstruct. 2018, 122, 631–635. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the device structures: (a) HT–VDMOS; (b) H–VDMOS; (c) C–VDMOS.
Figure 1. Schematic illustration of the device structures: (a) HT–VDMOS; (b) H–VDMOS; (c) C–VDMOS.
Crystals 13 00778 g001
Figure 2. Energy band schematics of the Si/SiC heterojunctions in (a) H–VDMOS and (b) HT–VDMOS.
Figure 2. Energy band schematics of the Si/SiC heterojunctions in (a) H–VDMOS and (b) HT–VDMOS.
Crystals 13 00778 g002
Figure 3. (a) Simulation results for the EC of the heterojunction with different NETL values at TETL = 0.1 μm. In the direction of the arrow, NETL is 0, 1 × 1017, 2 × 1017, 3 × 1017, 4 × 1017, 5 × 1017, 6 × 1017, 7 × 1017, 8 × 1017, 9 × 1017, and 1 × 1018 cm−3, respectively. (b) Simulation results for the EC of the heterojunction with different TETL values at NETL = 5 × 1017 cm−3. In the direction of the arrow, TETL is 0, 0.01, 0.02, 0.03, 0.04, 0.05, 0.06, 0.07, 0.08, 0.09 and 0.1 μm, respectively. The region corresponding to the X axes is indicated in Figure 1.
Figure 3. (a) Simulation results for the EC of the heterojunction with different NETL values at TETL = 0.1 μm. In the direction of the arrow, NETL is 0, 1 × 1017, 2 × 1017, 3 × 1017, 4 × 1017, 5 × 1017, 6 × 1017, 7 × 1017, 8 × 1017, 9 × 1017, and 1 × 1018 cm−3, respectively. (b) Simulation results for the EC of the heterojunction with different TETL values at NETL = 5 × 1017 cm−3. In the direction of the arrow, TETL is 0, 0.01, 0.02, 0.03, 0.04, 0.05, 0.06, 0.07, 0.08, 0.09 and 0.1 μm, respectively. The region corresponding to the X axes is indicated in Figure 1.
Crystals 13 00778 g003
Figure 4. Energy band schematics of the Si/SiC heterojunctions and the trap–assisted tunneling effect.
Figure 4. Energy band schematics of the Si/SiC heterojunctions and the trap–assisted tunneling effect.
Crystals 13 00778 g004
Figure 5. The trade–offs among the RON, EOX–M and ESi with the variation of (a) NETL, (b) TETL, (c) LETL and (d) WJFET. NETL = 5 × 1017 cm−3, TETL = 0.07 μm, the LETL = 1 μm and WJFET = 2μm are chosen as the optimized values. In each graph, the optimized values of the other three are utilized.
Figure 5. The trade–offs among the RON, EOX–M and ESi with the variation of (a) NETL, (b) TETL, (c) LETL and (d) WJFET. NETL = 5 × 1017 cm−3, TETL = 0.07 μm, the LETL = 1 μm and WJFET = 2μm are chosen as the optimized values. In each graph, the optimized values of the other three are utilized.
Crystals 13 00778 g005
Figure 6. The output I–V curves with VG values of 10 V, 15 V and 20 V and the reverse characteristics of the HT–VDMOS, H–VDMOS and C–VDMOS. The inset graph shows the variation of the RON of these three devices with the different VG values.
Figure 6. The output I–V curves with VG values of 10 V, 15 V and 20 V and the reverse characteristics of the HT–VDMOS, H–VDMOS and C–VDMOS. The inset graph shows the variation of the RON of these three devices with the different VG values.
Crystals 13 00778 g006
Figure 7. Simulation results of the Crss–VD curves with the test frequency value of 1 MHz.
Figure 7. Simulation results of the Crss–VD curves with the test frequency value of 1 MHz.
Crystals 13 00778 g007
Figure 8. Gate charge tests with the test circuit shown in the inset. The area of each device was set to 1 cm2. QGD values of 54 nC/cm2, 78 nC/cm2 and 63 nC/cm2 and QG values of 371 nC/cm2, 408 nC/cm2 and 365 nC/cm2 were obtained for the HT–VDMOS, H–VDMOS and C–VDMOS devices, respectively.
Figure 8. Gate charge tests with the test circuit shown in the inset. The area of each device was set to 1 cm2. QGD values of 54 nC/cm2, 78 nC/cm2 and 63 nC/cm2 and QG values of 371 nC/cm2, 408 nC/cm2 and 365 nC/cm2 were obtained for the HT–VDMOS, H–VDMOS and C–VDMOS devices, respectively.
Crystals 13 00778 g008
Figure 9. (a) Components of the RON corresponding to the geometrical positions in the structure. (b) Values of the components of the RON of the HT–VDMOS and C–VDMOS under different VG values.
Figure 9. (a) Components of the RON corresponding to the geometrical positions in the structure. (b) Values of the components of the RON of the HT–VDMOS and C–VDMOS under different VG values.
Crystals 13 00778 g009
Figure 10. (a) The influence of the QOX value on the RON of the HT–VDMOS and C–VDMOS. (b) The influence of the μ C H value on the RON of the HT–VDMOS.
Figure 10. (a) The influence of the QOX value on the RON of the HT–VDMOS and C–VDMOS. (b) The influence of the μ C H value on the RON of the HT–VDMOS.
Crystals 13 00778 g010
Figure 11. (a) Output I–V characteristic curves of the HT–VDMOS for different QI values. The inset graph shows the corresponding RON values. (b) Values of the components of the RON of the HT–VDMOS with different QI values. The VG is set to 15 V.
Figure 11. (a) Output I–V characteristic curves of the HT–VDMOS for different QI values. The inset graph shows the corresponding RON values. (b) Values of the components of the RON of the HT–VDMOS with different QI values. The VG is set to 15 V.
Crystals 13 00778 g011
Figure 12. Gate charge tests of the HT–VDMOS for different QI values. The area of each device was set to 1 cm2. The inset graph shows the corresponding QGD values.
Figure 12. Gate charge tests of the HT–VDMOS for different QI values. The area of each device was set to 1 cm2. The inset graph shows the corresponding QGD values.
Crystals 13 00778 g012
Table 1. Simulation Parameters of the mobility of 4H–SiC.
Table 1. Simulation Parameters of the mobility of 4H–SiC.
ParametersValueUnitParametersValueUnitParametersValueUnitParametersValueUnit
bn.cvt4.776 × 107cm2/Vsen.cvt1mun947cm2/Vsvsatn2 × 107cm/s
cn.cvt4.431 × 105cm2/Vsfeln.cvt1 × 1070cm2/Vstmun2betan2
taun.cvt0.0284kn.cvt2mup124cm2/Vsvsatp1.2 × 107cm/s
dn.cvt0.3333deln.cvt7.3 × 1012cm2/Vstmup2betap1
Table 2. Characteristics of HT–VDMOS and C–VDMOS.
Table 2. Characteristics of HT–VDMOS and C–VDMOS.
ParametersHT–VDMOSH–VDMOSC–VDMOSConditionsUnit
Breakdown voltage, BV152915221522ID = 0.1 μA/μm2V
On–state resistance, RON3.8830.7612.83VG = 10 V, VD = 1 VmΩ × cm2
3.6730.6210.72VG = 15 V, VD = 1 VmΩ × cm2
3.6030.549.99VG = 20 V, VD = 1 VmΩ × cm2
Miller capacitance, Crss306846F = 1 MHz, V = 400 VpF/cm2
Gate–drain charge, QGD547863VD = 400 V, ID = 300 AnC/cm2
Gate charge, QG371408365VD = 400 V, ID = 300 AnC/cm2
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

Chen, H.; Zhang, Y.; Zhou, R.; Wang, Z.; Lu, C.; Li, Z.; Zhang, B. A Novel Low On–State Resistance Si/4H–SiC Heterojunction VDMOS with Electron Tunneling Layer Based on a Discussion of the Hetero–Transfer Mechanism. Crystals 2023, 13, 778. https://doi.org/10.3390/cryst13050778

AMA Style

Chen H, Zhang Y, Zhou R, Wang Z, Lu C, Li Z, Zhang B. A Novel Low On–State Resistance Si/4H–SiC Heterojunction VDMOS with Electron Tunneling Layer Based on a Discussion of the Hetero–Transfer Mechanism. Crystals. 2023; 13(5):778. https://doi.org/10.3390/cryst13050778

Chicago/Turabian Style

Chen, Hang, Yourun Zhang, Rong Zhou, Zhi Wang, Chao Lu, Zehong Li, and Bo Zhang. 2023. "A Novel Low On–State Resistance Si/4H–SiC Heterojunction VDMOS with Electron Tunneling Layer Based on a Discussion of the Hetero–Transfer Mechanism" Crystals 13, no. 5: 778. https://doi.org/10.3390/cryst13050778

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