Next Article in Journal
Electric Discharge Machining on Stainless Steel Using a Blend of Copper and Fly Ash as the Electrode Material
Next Article in Special Issue
Novel Photonic Applications of Silicon Carbide
Previous Article in Journal
On the Electronic Structure of 2H-MoS2: Correlating DFT Calculations and In-Situ Mechanical Bending on TEM
Previous Article in Special Issue
A New Cell Topology for 4H-SiC Planar Power MOSFETs for High-Frequency Switching
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of Carrier Mobilities in 4H-SiC MOSFETS Using Hall Analysis

1
Department of Physics, Auburn University, Auburn, AL 36849, USA
2
SemiQ Inc., Lake Forest, CA 92630, USA
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(19), 6736; https://doi.org/10.3390/ma15196736
Submission received: 2 September 2022 / Revised: 22 September 2022 / Accepted: 25 September 2022 / Published: 28 September 2022
(This article belongs to the Special Issue Silicon Carbide: Material Growth, Device Processing and Applications)

Abstract

:
The channel conduction in 4H-SiC metal–oxide–semiconductor field effect transistors (MOSFETs) are highly impacted by charge trapping and scattering at the interface. Even though nitridation reduces the interface trap density, scattering still plays a crucial role in increasing the channel resistance in these transistors. In this work, the dominant scattering mechanisms are distinguished for inversion layer electrons and holes using temperature and body-bias-dependent Hall measurements on nitrided lateral 4H-SiC MOSFETs. The effect of the transverse electric field ( E e f f ) on carrier mobility is analyzed under strong inversion condition where surface roughness scattering becomes prevalent. Power law dependencies of the electron and hole Hall mobility for surface roughness scattering are determined to be E e f f 1.8 and E e f f 2.4 , respectively, analogous to those of silicon MOSFETs. Moreover, for n-channel MOSFETs, the effect of phonon scattering is observed at zero body bias, whereas in p-channel MOSFETs, it is observed only under negative body biases. Along with the identification of regimes governed by different scattering mechanisms, these results highlight the importance of the selection of substrate doping and of E e f f in controlling the value of channel mobility in 4H-SiC MOSFETs.

1. Introduction

Silicon carbide (4H-SiC) is one of the primary wide-band-gap semiconductors for high power and harsh environment applications because of its physical properties, such as a high critical electric field and high thermal conductivity [1]. Discrete 4H-SiC diodes and metal–oxide–semiconductor field effect transistors (MOSFETs) are being widely adopted for high voltage power conversion in hybrid/electric vehicles [2], solar and wind energy generation [3], and various high temperature applications [4], enabling significant advances for next-generation energy efficient power systems. Moreover, 4H-SiC is a unique candidate for the development of an integrated circuit (IC) technology that is still in its infancy. 4H-SiC IC technology operating at very high temperatures >300 °C [5,6] is very attractive, as it enables operation in environments and ambientes that are not accessible to conventional silicon or silicon on insulator (SOI) platforms [7,8]. Lateral complementary metal–oxide–semiconductor (CMOS) IC technology in 4H-SiC is desirable for the fabrication of large-scale integration devices due to its high noise immunity and low static power consumption [9]. To be materialized, such technology demands both n- and p-channel MOSFETs capable of operating at high-temperatures. While n-channel MOSFETs have reached a suitable degree of maturity through the nitridation of the SiO2/4H-SiC interface [10,11,12], mechanisms governing electrical transport in their p-channel counterparts must be investigated to reach comparable levels of development in terms of channel conductivity and device stability.
Recent studies [10,13] on 4H-SiC MOSFETs use field-effect mobility models to analyze channel transport. Additionally, research by Mikami et al. [13] explored the importance of body bias and explained how a body bias experiment can provide predictions of transistor characteristics for MOSFETs made with higher or lower substrate doping using field effect mobility on the weak inversion region. However, the channel scattering mechanisms for p-channel 4H-SiC MOSFET remain unexplored using Hall analysis. Hall mobility is more accurate than field effect mobility, as the carrier concentration is measured independently and does not get affected by trapped charges. A comprehensive study for both kinds of carriers will enable control of the operation region and the preferable substrate doping concentrations for high-temperature IC design. To this end, the channel mobility of nitric oxide (NO)-annealed n- and p-channel 4H-SiC MOSFETs in the strong inversion regime was characterized using Hall measurements. Measurements were performed using the body-bias technique, which allows variation of the effective electric field ( E e f f ) in the channel [13,14]. This approach allows one to differentiate and establish the dominant scattering mechanisms limiting carrier transport, in different carrier density and substrate doping regimes. The power law that governs the relation between μ H a l l and E e f f was determined for both electrons and holes in the surface-roughness-scattering-dominant regime. In addition, it is demonstrated that a combination of temperature and body bias enables observation of the dominant mechanisms transitioning from surface roughness to phonon scattering.

2. Materials and Methods

2.1. Fundamentals on Channel-Transverse Electric Fields and Mobility

The channel’s resistance depends on the channel’s inversion-carrier concentration and its mobility ( μ H a l l ), which are related to the atomic-scale composition, structure, and defects at the SiO2/4H-SiC interface. The mobile carrier’s concentration is reduced by traps at electrically-active defect sites, and the channel’s mobility is degraded by the interfacial scattering processes. When the MOSFFET is on, the dominant scattering mechanisms at a SiO2/4H-SiC interface that limits μ H a l l are: (i) Coulomb scattering occurring due to trapped charges present at or near the interface oxide and from ionized impurities in the depletion region close to the channel; (ii) phonon scattering occurring due to the interaction of the carriers with lattice vibration in the channel and the surface; and (iii) surface roughness due to imperfections of the 4H-SiC interface [15,16,17]. The resulting carrier mobility is given by Matthiesen’s rule [18]:
1 μ H a l l = 1 μ C   + 1 μ p h + 1 μ S R  
Here, μ C , μ p h , and μ S R are the Coulomb, phonon, and surface roughness scattering limited mobilities, respectively.
On Si MOSFETs [19,20], earlier studies suggest that the transverse electric field E e f f is an important parameter for influencing scattering in the channel. The magnitude of E e f f can be obtained as [19],
E e f f = 1 ϵ 2 N A q ϵ . 2 ϕ B ± V B S + η q n s  
The first term in Equation (2) represents the amount of depletion charge, and the second term stands for the induced inversion layer charge excluding trapped charges. Here the “+” sign is applicable for n-channel MOSFETs and the “ “ sign is for p-channel MOSFETs. A schematic diagram of these two charge distributions is shown in the left panel of Figure 1. In Equation (2), η is a constant with a value of 0.5 for n-channel and 0.33 for p-channel Si MOSFETs at 27   ° C [19,20,21,22], which are considered the same for 4H-SiC [13,14,23,24], but this choice does not affect the results and may need further studies in a wider effective field range. ϵ is the dielectric constant, q is the electron charge, n s is the free carrier concentration, and NA is the p-well substrate doping concentration for n-channel MOSFETs. VBS is the substrate or body to source bias. ϕ B = K T q   ln N A n i is the bulk potential. K is the Boltzmann constant and n i denotes the temperature-dependent intrinsic carrier concentration. Note, for p-channel MOSFET, the n-well substrate doping concentration is denoted by ND. Application of body/substrate bias acts as a separate knob to control the depletion width without affecting the channel carrier concentration. A variation of channel electric field to study its effect on the conduction and the effects of substrate doping concentration is possible through VBS.

2.2. Experimental Methods

Hall bar MOSFETs with a channel length of 600 μm and a channel width of 60 μm were fabricated on a 4° off-axis (0001), Si-face oriented 4H-SiC substrate with a p-type epilayer (NAND = 6 × 1015 cm−3) for n-channel MOSFETs and an n-type epilayer (NDNA = 6.2 × 1015 cm−3) for p-channel MOSFETs. Additional to the Hall bar MOSFETs, T-gated MOSFETs were also present in the chip with a 200 μm × 200 μm channel (width, length). The cross-sectional schematic of the Hall bar device is shown in Figure 1 (left). A nominally uniform doping profile was created using nitrogen and aluminum ion implantations at 700 °C to form n+ and p+ layers on the MOSFETs, respectively, to form source, drain and body contacts. The implanted layers were then activated by annealing at 1650 °C for 30 min in flowing Ar with a graphitic cap to protect the surface. After removal of the cap, gate oxidation was carried out at 1150 °C for 10 h followed by post oxidation annealing at 1175 °C for 2 h in NO. The thickness of the gate oxide was measured to be about ~60 nm and 55 nm by the capacitance voltage method (CV) for n- and p-channel MOSFETs, respectively. Contact photolithography and reactive ion etching were carried out to remove the oxide from the source/drain/body (S/D/B) regions and Hall voltage (Vh) terminals. For n-channel MOSFET, Al was evaporated for all the contacts and annealed at 800 °C for 30 min in flowing Ar. On the other hand, for p-channel MOSFET, Ti was sputtered and lifted off to define the S/D/B/Vh contacts, as shown in Figure 1 (right), following which annealing at 1000 °C for 2 min was performed in flowing Ar to form ohmic contacts. Subsequently, aluminum (55 nm) was thermally evaporated as the gate metal and patterned using lift-off. An overlayer of Au/Cr was sputtered on all the contacts of both kinds of MOSFETs. Finally, the Hall bar MOSFET was mounted on a ceramic chip with epoxy and wire bonded to gold pads on the chip.

3. Results and Discussions

3.1. Classification of the Dominant Scattering Mechanisms

After fabrication of the 4H-SiC MOSFETs, transfer characteristics were obtained at various temperatures. The temperature for the n-channel MOSFETs was varied from 77 to 373 K. For the p-channel MOSFETs, on the other hand, the temperature was raised above room temperature (300 K) until 573 K. The absolute value of threshold voltage became smaller at higher temperatures, as shown in Figure 2. This is associated with the reduction of occupied interface trap densities (Dit) at higher temperatures [25]. In addition, the field effect mobility for electrons ( μ f e ,   n ) was observed to increase from 77 to 296 K. For the p-channel MOSFET, the field effect mobility of holes was observed to be weakly dependent on temperature under strong inversion.
To confirm the dominant scattering mechanisms in these devices, Hall measurements were carried out under a perpendicular magnetic field of 0.6 T. At a high gate voltage overdrive, the Hall and field effect mobilities merge, and at low and intermediate gate voltages, field effect mobilities are lower than Hall mobilities due to the presence of trapped charges (see supplementary material). Figure 3 shows μ H a l l versus carrier concentration curves at different temperatures and at zero body bias for electrons and holes in n- and p-channel 4H-SiC MOSFETs. To obtain μ H a l l versus carrier concentration plot, the gate to source voltage V g was varied from +2 to +10 V for n-channel and −6 to −20 V for p-channel MOSFETs at constant drain to source voltage V D S = 0.75   V . The threshold voltage was found to be +5 V for electrons (at 293 K) and −10 V for holes (300 K), based on linear extrapolation of the ID–VG characteristics (Figure 2). For n-channel MOSFETs, Hall mobility above room temperature drops. An increase in mobility from 77 to 293 K is a signature of prevalent Coulomb scattering, whereas above 293 K, mobility decreases with increasing temperature when phonon scattering is dominant. Therefore, for electrons, transport is limited by Coulomb and phonon scattering. From the trend of the experimental results in figure in Section 3.3 and other published results [23,26], electron mobility will reduce below 77 K due to increased Coulomb scattering. Above 373 K, mobility would also decrease, but due to increased phonon scattering. The presence of surface roughness scattering at higher concentration can also be observed in earlier reports [14]. Conversely, for holes in p-channel MOSFET, Hall mobility is seen to increase at lower carrier concentration (for approximately p s < 6 × 10 11   cm 2 at 300 K), signifying dominant Coulomb scattering, and becomes independent of temperature at approximately p s = 10 12   cm 2 (see supplementary materials). Temperature-independent mobility is a sign of surface roughness scattering. Therefore, hole mobility in 4H-SiC p-channel MOSFET is limited by Coulomb scattering at weak inversion and dominant surface roughness scattering at strong inversion.
After distinguishing the dominant scattering mechanisms in these devices, Hall measurements were carried out under the effect of body bias (VBS) so that the Hall mobility ( μ H a l l ) can be studied as a function of transverse electric field E e f f in the strong inversion regime. Recent studies [13,14] observed the effect of transverse electric field on mobility in the strong inversion regime maintaining a fixed carrier concentration. A constant carrier concentration realizes fixed screening from scattering centers and enables the assessment of the sole effect of surface roughness scattering on mobility.

3.2. Analysis of Surface Roughness Scattering Using Body Bias Measurements

First, to confirm the body bias effect on each transistor, a transfer characteristic at different body bias levels was measured, as shown in Figure 4. The threshold voltages (Vth) at 300 K and VBS = 0 were approximately +4 V and −10 V for n- and p-channel MOSFETs respectively, based on linear extrapolation of the ID–VG characteristics. It can be observed that the threshold voltage increases as the VBS goes from forward to reverse bias for both types of MOSFETs.
A positive increment in the VBS widens the depletion layer between the source and the body junction, thereby requiring a higher gate voltage to achieve same level of drain current (Id) as in lower VBS. The plot of μ H a l l as a function of carrier concentration in the strong inversion region for different body biases is shown in Figure 5. An increase in reverse body bias increases the magnitude of the transverse electric field (Equation (2)) and pushes the carriers closer to the surface, making the channel width thinner (~1 nm) at strong inversion.
Hence, at a high gate voltage, when mobile holes are closer to the surface, surface roughness scattering becomes even more prevalent through the influence of the varying perturbed potential energy [15]. This is attributed to charge carrier wave functions becoming more susceptible to the fluctuating perturbation near the interface stemming from the crystal miscut (4° off axis substrate) and nano-steps/roughness. As a result, an increased E e f f lowers the value of μ H a l l . Conversely, a negative body bias decreases the value of E e f f , which widens the channel region and lessens the impact of the surface roughness scattering.
These results were used to extract the functional dependence of μ H a l l at a fixed carrier concentration to analyze surface roughness scattering under constant screening. For this purpose, using Equation (2), n s   and   p s were converted to E e f f for each VBS above 0 V in the strong inversion regime. Next, μ H a l l was extracted at fixed carrier concentrations ( 1.0 × 10 12 , 1.5 × 10 12 ,   2.0 × 10 12   and   2.5 × 10 12 cm−2) and plotted against E e f f , as shown in Figure 6. The values of VBS were chosen above zero volts, where the surface roughness becomes the dominant mechanism and there is a minimal contribution from phonon scattering. The curves have been fit using a power law function of E e f f , yielding exponents of 1.8 ± 0.2 and 2.4 ± 0.3 for channel electrons and holes in 4H-SiC n- and p-channel MOSFETs, respectively. These values are close to those found earlier for surface roughness scattering in Si MOSFETs [20], suggesting that at high E e f f , surface imperfections have a similar effect on channel carriers in 4H-SiC as in Si MOSFETs. However, differently from Si, where μ H a l l versus E e f f curves merge with each other when surface roughness scattering becomes the dominant regime, this universal behavior is not visible for 4H-SiC in Figure 6. This is consistent with earlier studies on 4H-SiC n-channel MOSFETs [14,23,24,27,28]. Perhaps for a single 4H-SiC MOSFET, the measurable range of E e f f is not large enough before the breakdown of the gate dielectric to observe the merging of μ H a l l at higher E e f f , and a larger range of measurements using devices with different substrate doping is necessary, as reported in [24].

3.3. Phonon-Scattering-Limited Mobility

For n-channel MOSFETs, dropping mobility was observed above room temperature at zero body bias, which signifies the dominance of phonon-scattering-limited mobility, as seen in Figure 7. This fact is consistent with recent work [23] which reported that phonon-scattering-limited mobility can be observed for n-channel MOSFETs fabricated on lightly doped ( 5 × 10 15 cm−3) p-type epitaxial layers. However, for p-channel MOSFETs, phonon-scattering-limited mobility was absent at zero body bias in the measured temperature range. In this case, phonon scattering was visible only when the source to body junction was kept at a forward bias (VBS < 0). Then, the power law-dependence of the μ H a l l on E e f f was observed to give an exponent of −0.32 (see supplementary materials). A power law of E e f f 1 / 3 is an indication of phonon scattering [20,23]. In our study, a negative body bias was used to replicate a lightly doped n-type substrate through the “effective doping concentration” [14] of the n-well N D , e f f , given by:
N D , e f f = N D 1 + V B S 2 ϕ B .
At a body bias V B S = 1.3   V and a fixed p s = 10 12 cm−2, a low value of N D , e f f (approximately 3 × 10 15 cm−3) can be maintained in the 300 to 498 K (27 °C to 225 °C) temperature range. Figure 7 shows that μ H a l l increases from room temperature to 398 K due to Coulomb scattering, after which it decreases owing to the dominance of phonon scattering following a power law of T 0.9 .
The power of T found here is near to the theoretical value of T−1 [29]. Under negative V B S , the channel is thick enough to interact with the n-well lattice vibrations at T > 398 K, and the hole mobility is limited by phonon scattering. This is consistent with n-channel 4H-SiC MOSFETs fabricated on lightly doped substrates, where phonon-scattering-limited mobility at low E e f f or N D , e f f can be observed [23]. Conversely, for V B S = 0 V, channel thickness is comparatively thinner than at V B S = 1.3   V . In this case, no photon-limited scattering is observable, but μ H a l l is limited by the cumulative effect of Coulomb and surface roughness scattering. As a result, μ H a l l increases slightly at higher temperatures.

4. Conclusions

In conclusion, the dominant scattering mechanisms in n- and p- 4H-SiC MOSFET channels were distinguished using Hall measurements. Electron mobility in n-channel MOSFET is limited by Coulomb scattering in weak inversion and a combination of phonon and surface roughness scattering in the strong inversion region. On the other hand, hole mobility in p-channel MOSFETs is limited primarily by Coulomb and surface roughness scattering. Body-bias- and temperature-dependent Hall measurements were performed and analyzed under strong inversion to study mobility in terms of effective transvers electric field ( E e f f ). As body bias is increased, a higher value of E e f f confines the mobile carriers near the 4H-SiC/SiO2 interface, resulting in surface-roughness-scattering-limited mobility. A changing body bias was applied to determine the power-law dependence of channel electron and hole mobility on the E e f f for surface roughness scattering in 4H-SiC as E e f f 1.8 and E e f f 2.4 , respectively. Furthermore, with the application of negative VBS, the depletion layer between the source and body contact decreases, and the resulting low E e f f leads to higher mobility at a given gate bias. For p-channel MOSFET, at a forward VBS (−1.3 V at 398 K), the channel is thick enough to interact with the n-well lattice vibration, giving rise to a phonon-scattering-limited hole mobility. Therefore, at low negative VBS, the channel conduction behaves like a MOSFET with a lightly doped substrate. These new findings emphasize the significance of substrate doping selection for 4H-SiC MOSFETs. Additionally, for the development of highly efficient 4H-SiC CMOS devices for high temperature operation, the three types of scattering processes identified here must be considered to depend on the value of the transverse electric field in the channel and the operating temperature.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ma15196736/s1. Figure S1: Field effect versus Hall mobility. Figure S2: Surface roughness scattering at a high hole concentration. Figure S3: Hole mobility versus transverse electric field for different body biases at room temperature.

Author Contributions

The research was conducted under the supervision of S.D. (Sarit Dhar). The devices were fabricated and characterized by S.D. (Suman Das) and Y.Z. The data were analyzed, and the manuscript was prepared by S.D. (Suman Das). A.A. helped in fabrication and measurements. M.A.K. and S.D. (Sarit Dhar) reviewed and improved the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially funded by the US Department of Energy (subcontracted to National Renewable Energy Laboratory, grant NREL-AHL-9-92632-01).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

Acknowledgments

The authors acknowledge the support of Tamara Isaacs-Smith of Auburn University for help with device fabrication.

Conflicts of Interest

The authors do not have any potential conflict of interest to disclose.

References

  1. Baliga, B.J. Fundamentals of Power Semiconductor Devices; Springer Science & Business Media: Berlin, Germany, 2010; ISBN 978-0-387-47314-7. [Google Scholar]
  2. Agarwal, A.K. An Overview of SiC Power Devices. In Proceedings of the 2010 International Conference on Power, Control and Embedded Systems, Allahabad, India, 29 November–1 December 2010; pp. 1–4. [Google Scholar]
  3. He, J.; Zhao, T.; Jing, X.; Demerdash, N.A.O. Application of Wide Bandgap Devices in Renewable Energy Systems—Benefits and Challenges. In Proceedings of the 2014 International Conference on Renewable Energy Research and Application (ICRERA), Milwaukee, WI, USA, 19–22 October 2014; pp. 749–754. [Google Scholar]
  4. Hunter, G.; Neudeck, P.; Xu, J.C.; Lukco, D.; Trunek, A.J.; Artale, M.; Lampard, P.; Androjna, D.; Makel, D.; Ward, B.; et al. Development of SiC-Based Gas Sensors for Aerospace Applications. MRS Online Proc. Libr. 2004, 815. [Google Scholar] [CrossRef]
  5. Rahman, A.; Roy, S.; Murphree, R.; Kotecha, R.; Addington, K.; Abbasi, A.; Mantooth, H.A.; Francis, A.M.; Holmes, J.; Di, J. High-Temperature SiC CMOS Comparator and Op-Amp for Protection Circuits in Voltage Regulators and Switch-Mode Converters. IEEE J. Emerg. Sel. Top. Power Electron. 2016, 4, 935–945. [Google Scholar] [CrossRef]
  6. Holmes, J.; Francis, A.M.; Getreu, I.; Barlow, M.; Abbasi, A.; Mantooth, H.A. Extended High-Temperature Operation of Silicon Carbide CMOS Circuits for Venus Surface Application. J. Microelectron. Electron. Packag. 2016, 13, 143–154. [Google Scholar] [CrossRef]
  7. Kappert, H.; Kordas, N.; Dreiner, S.; Paschen, U.; Kokozinski, R. High Temperature SOI CMOS Technology and Circuit Realization for Applications up to 300 °C. In Proceedings of the 2015 IEEE International Symposium on Circuits and Systems (ISCAS), Lisbon, Portugal, 24–27 May 2015. [Google Scholar] [CrossRef]
  8. Homberger, J.M.; Cilio, E.; Schupbach, R.M.; Lostetter, A.B.; Mantooth, H.A. A High-Temperature Multichip Power Module (MCPM) Inverter Utilizing Silicon Carbide (SiC) and Silicon on Insulator (SOI) Electronics. In Proceedings of the 2006 37th IEEE Power Electronics Specialists Conference, Jeju, Korea, 18–22 June 2006; pp. 1–7. [Google Scholar]
  9. Liu, D.; Svensson, C. Power Consumption Estimation in CMOS VLSI Chips. IEEE J. Solid-State Circuits 1994, 29, 663–670. [Google Scholar] [CrossRef]
  10. Tachiki, K.; Kimoto, T. Improvement of Both N- and p-Channel Mobilities in 4H-SiC MOSFETs by High-Temperature N₂ Annealing. IEEE Trans. Electron Devices 2021, 68, 638–644. [Google Scholar] [CrossRef]
  11. Tachiki, K.; Kaneko, M.; Kimoto, T. Mobility Improvement of 4H-SiC (0001) MOSFETs by a Three-Step Process of H2 Etching, SiO2 Deposition, and Interface Nitridation. Appl. Phys. Express 2021, 14, 031001. [Google Scholar] [CrossRef]
  12. Tachiki, K.; Mikami, K.; Ito, K.; Kaneko, M.; Kimoto, T. Mobility Enhancement in Heavily Doped 4H-SiC (0001), ( 11 2 ¯ 0 ), and ( 1 1 ¯ 00 ) MOSFETs via an Oxidation-Minimizing Process. Appl. Phys. Express 2022, 15, 071001. [Google Scholar] [CrossRef]
  13. Mikami, K.; Tachiki, K.; Ito, K.; Kimoto, T. Body Doping Dependence of Field-Effect Mobility in Both n- and p-Channel 4H-SiC Metal-Oxide-Semiconductor Field-Effect Transistors with Nitrided Gate Oxides. Appl. Phys. Express 2022, 15, 036503. [Google Scholar] [CrossRef]
  14. Hirai, H.; Hatakeyama, T.; Sometani, M.; Okamoto, M.; Harada, S.; Okumura, H.; Yamaguchi, H. Difference in Electron Mobility at 4H–SiC/SiO2 Interfaces with Various Crystal Faces Originating from Effective-Field-Dependent Scattering. Appl. Phys. Lett. 2020, 117, 042101. [Google Scholar] [CrossRef]
  15. Yamanaka, T.; Fang, S.J.; Lin, H.-C.; Snyder, J.P.; Helms, C.R. Correlation between Inversion Layer Mobility and Surface Roughness Measured by AFM. IEEE Electron Device Lett. 1996, 17, 178–180. [Google Scholar] [CrossRef]
  16. Pirovano, A.; Lacaita, A.L.; Ghidini, G.; Tallarida, G. On the Correlation between Surface Roughness and Inversion Layer Mobility in Si-MOSFETs. IEEE Electron Device Lett. 2000, 21, 34–36. [Google Scholar] [CrossRef]
  17. Matsumoto, Y.; Uemura, Y. Scattering Mechanism and Low Temperature Mobility of MOS Inversion Layers. Jpn. J. Appl. Phys. 1974, 13, 367. [Google Scholar] [CrossRef]
  18. Sze, S.M.; Ng, K.K. Physics of Semiconductor Devices, 3rd ed.; Wiley-Interscience: Hoboken, NJ, USA, 2007; ISBN 978-0-471-14323-9. [Google Scholar]
  19. Sun, S.C.; Plummer, J.D. Electron Mobility in Inversion and Accumulation Layers on Thermally Oxidized Silicon Surfaces. IEEE Trans. Electron Devices 1980, 27, 1497–1508. [Google Scholar] [CrossRef]
  20. Takagi, S.; Toriumi, A.; Iwase, M.; Tango, H. On the Universality of Inversion Layer Mobility in Si MOSFET’s: Part I-Effects of Substrate Impurity Concentration. IEEE Trans. Electron Devices 1994, 41, 2357–2362. [Google Scholar] [CrossRef]
  21. Arora, N.D.; Gildenblat, G.S. A Semi-Empirical Model of the MOSFET Inversion Layer Mobility for Low-Temperature Operation. IEEE Trans. Electron Devices 1987, 34, 89–93. [Google Scholar] [CrossRef]
  22. Sabnis, A.G.; Clemens, J.T. Characterization of the Electron Mobility in the Inverted <100> Si Surface. In Proceedings of the 1979 International Electron Devices Meeting, Washington, DC, USA, 3–5 December 1979; pp. 18–21. [Google Scholar]
  23. Noguchi, M.; Iwamatsu, T.; Amishiro, H.; Watanabe, H.; Miura, N.; Kita, K.; Yamakawa, S. Carrier Transport Properties in Inversion Layer of Si-Face 4H–SiC MOSFET with Nitrided Oxide. Jpn. J. Appl. Phys. 2019, 58, 031004. [Google Scholar] [CrossRef]
  24. Noguchi, M.; Watanabe, T.; Watanabe, H.; Kita, K.; Miura, N. Comparative Study of Hall Effect Mobility in Inversion Layer of 4H-SiC MOSFETs With Nitrided and Phosphorus-Doped Gate Oxides. IEEE Trans. Electron Devices 2021, 68, 6321–6329. [Google Scholar] [CrossRef]
  25. Das, S.; Isaacs-Smith, T.; Ahyi, A.; Kuroda, M.A.; Dhar, S. High Temperature Characteristics of Nitric Oxide Annealed P-Channel 4H-SiC Metal Oxide Semiconductor Field Effect Transistors. J. Appl. Phys. 2021, 130, 225701. [Google Scholar] [CrossRef]
  26. Noguchi, M.; Iwamatsu, T.; Amishiro, H.; Watanabe, H.; Kita, K.; Miura, N. Coulomb-Limited Mobility in 4H-SiC MOS Inversion Layer as a Function of Inversion-Carrier Average Distance from MOS Interface. Jpn. J. Appl. Phys. 2020, 59, 051006. [Google Scholar] [CrossRef]
  27. Uhnevionak, V.; Burenkov, A.; Strenger, C.; Ortiz, G.; Bedel-Pereira, E.; Mortet, V.; Cristiano, F.; Bauer, A.J.; Pichler, P. Comprehensive Study of the Electron Scattering Mechanisms in 4H-SiC MOSFETs. IEEE Trans. Electron Devices 2015, 62, 2562–2570. [Google Scholar] [CrossRef]
  28. Potbhare, S.; Goldsman, N.; Lelis, A.; McGarrity, J.M.; McLean, F.B.; Habersat, D. A Physical Model of High Temperature 4H-SiC MOSFETs. IEEE Trans. Electron Devices 2008, 55, 2029–2040. [Google Scholar] [CrossRef]
  29. Takagi, S.; Toriumi, A.; Iwase, M.; Tango, H. On the Universality of Inversion Layer Mobility in Si MOSFET’s: Part II-Effects of Surface Orientation. IEEE Trans. Electron Devices 1994, 41, 2363–2368. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (left) Schematic diagram of lateral n-channel Hall MOSFET with the transverse electric field E e f f shown, which depends on space charge and inversion charge in the channel at a gate voltage higher than the threshold voltage. (right) Top view of a schematic of the fabricated Hall bar MOSFET. Vh1, Vh2, Vh3, and Vh4 represent the Hall voltage terminals.
Figure 1. (left) Schematic diagram of lateral n-channel Hall MOSFET with the transverse electric field E e f f shown, which depends on space charge and inversion charge in the channel at a gate voltage higher than the threshold voltage. (right) Top view of a schematic of the fabricated Hall bar MOSFET. Vh1, Vh2, Vh3, and Vh4 represent the Hall voltage terminals.
Materials 15 06736 g001
Figure 2. Field effect mobility of electrons and holes versus gate voltage at different temperatures for n- and p-channel 4H-SiC MOSFETs. For n-channel MOSFETs, channel width/length was 60 µm/600 µm, oxide thickness was 60 nm, Vds was set to 0.4 V, and temperature was varied from 77 to 373 K. For p-channel MOSFETs, channel width/length was 200 µm/200 µm, oxide thickness was 55 nm, Vds was set to 0.1 V, and temperature was varied from 300 to 573 K.
Figure 2. Field effect mobility of electrons and holes versus gate voltage at different temperatures for n- and p-channel 4H-SiC MOSFETs. For n-channel MOSFETs, channel width/length was 60 µm/600 µm, oxide thickness was 60 nm, Vds was set to 0.4 V, and temperature was varied from 77 to 373 K. For p-channel MOSFETs, channel width/length was 200 µm/200 µm, oxide thickness was 55 nm, Vds was set to 0.1 V, and temperature was varied from 300 to 573 K.
Materials 15 06736 g002
Figure 3. Typical Hall mobility as a function of carrier concentration at different temperatures for n- and p-channel MOSFETs with channel width 60 µm and channel length 600 µm for both kinds of Hall bar MOSFET, and Vds was set to 0.75 V.
Figure 3. Typical Hall mobility as a function of carrier concentration at different temperatures for n- and p-channel MOSFETs with channel width 60 µm and channel length 600 µm for both kinds of Hall bar MOSFET, and Vds was set to 0.75 V.
Materials 15 06736 g003
Figure 4. Transfer characteristics at room temperature (300 K) for different body biases (VBS) for n- and p-channel MOSFETs. Id–Vg sweeps were taken at a constant drain-to-source voltage (0.75 V). A change in Vth can be observed due to the modulation of the depletion layer width at the body to source junction with the application of body bias. Note that the y-axis scales are different for the two kinds of MOSFETs.
Figure 4. Transfer characteristics at room temperature (300 K) for different body biases (VBS) for n- and p-channel MOSFETs. Id–Vg sweeps were taken at a constant drain-to-source voltage (0.75 V). A change in Vth can be observed due to the modulation of the depletion layer width at the body to source junction with the application of body bias. Note that the y-axis scales are different for the two kinds of MOSFETs.
Materials 15 06736 g004
Figure 5. Hall mobility as a function of channel carrier concentration at different body biases at 27 °C. The body bias was varied from +1 to −3 V for n-channel MOSFETs and −1 to +5 V for p-channel MOSFETs; the gate bias VGS was swept to modulate the carrier concentrations under a fixed body bias. A forward body bias value decreases the depletion layer width between source and body, increasing mobility, and at reverse body bias, Eeff increases due to expansion of depletion layer, which in turn decreases the channel mobility.
Figure 5. Hall mobility as a function of channel carrier concentration at different body biases at 27 °C. The body bias was varied from +1 to −3 V for n-channel MOSFETs and −1 to +5 V for p-channel MOSFETs; the gate bias VGS was swept to modulate the carrier concentrations under a fixed body bias. A forward body bias value decreases the depletion layer width between source and body, increasing mobility, and at reverse body bias, Eeff increases due to expansion of depletion layer, which in turn decreases the channel mobility.
Materials 15 06736 g005
Figure 6. Transverse electric field (Eeff) dependence of Hall mobility ( μ H a l l ) at different body biases (−1, −2, and −3 V for n-channel and 0, +1, +3, +5, and +7 V for p-channel) at room temperature (27 °C) for different carrier concentrations ( 1.0 × 10 12 , 1.5 × 10 12 ,   2.0 × 10 12 ,   and   2.5 × 10 12 cm−2), in log–log scale. The different Eeff points for a particular curve were obtained from different body biases at a fixed carrier concentration from Figure 5. The average power law that μ Hall follows on E e f f is 1.8 ± 0.2 for n-channel MOSFETs and 2.4 ± 0.3 for p-channel MOSFETs, providing the power of E e f f for surface roughness scattering for electrons and holes in 4H-SiC MOSFETs. Error bars signify ~5% error estimated based on variation in at least 3 measurements.
Figure 6. Transverse electric field (Eeff) dependence of Hall mobility ( μ H a l l ) at different body biases (−1, −2, and −3 V for n-channel and 0, +1, +3, +5, and +7 V for p-channel) at room temperature (27 °C) for different carrier concentrations ( 1.0 × 10 12 , 1.5 × 10 12 ,   2.0 × 10 12 ,   and   2.5 × 10 12 cm−2), in log–log scale. The different Eeff points for a particular curve were obtained from different body biases at a fixed carrier concentration from Figure 5. The average power law that μ Hall follows on E e f f is 1.8 ± 0.2 for n-channel MOSFETs and 2.4 ± 0.3 for p-channel MOSFETs, providing the power of E e f f for surface roughness scattering for electrons and holes in 4H-SiC MOSFETs. Error bars signify ~5% error estimated based on variation in at least 3 measurements.
Materials 15 06736 g006
Figure 7. Hall mobility ( μ H a l l ) versus temperature T for different body biases is plotted for a fixed n s   / or   p s = 10 12 cm−2. For n-channel MOSFET, Coulomb scattering is dominant below RT; above RT phonon scattering can be observed at zero body bias. However, for p-channel MOSFETs, μ Hall is limited by a combined effect of Coulomb and surface roughness scattering, VBS = 0 V. For VBS = −1.3 V, μ H a l l increases linearly until 398 K (125 °C) due to Coulomb scattering. At higher temperatures, phonon scattering dominates and μ Hall decreases with temperature as T−0.9. Error bars signify variation in in at least 3 measurements.
Figure 7. Hall mobility ( μ H a l l ) versus temperature T for different body biases is plotted for a fixed n s   / or   p s = 10 12 cm−2. For n-channel MOSFET, Coulomb scattering is dominant below RT; above RT phonon scattering can be observed at zero body bias. However, for p-channel MOSFETs, μ Hall is limited by a combined effect of Coulomb and surface roughness scattering, VBS = 0 V. For VBS = −1.3 V, μ H a l l increases linearly until 398 K (125 °C) due to Coulomb scattering. At higher temperatures, phonon scattering dominates and μ Hall decreases with temperature as T−0.9. Error bars signify variation in in at least 3 measurements.
Materials 15 06736 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Das, S.; Zheng, Y.; Ahyi, A.; Kuroda, M.A.; Dhar, S. Study of Carrier Mobilities in 4H-SiC MOSFETS Using Hall Analysis. Materials 2022, 15, 6736. https://doi.org/10.3390/ma15196736

AMA Style

Das S, Zheng Y, Ahyi A, Kuroda MA, Dhar S. Study of Carrier Mobilities in 4H-SiC MOSFETS Using Hall Analysis. Materials. 2022; 15(19):6736. https://doi.org/10.3390/ma15196736

Chicago/Turabian Style

Das, Suman, Yongju Zheng, Ayayi Ahyi, Marcelo A. Kuroda, and Sarit Dhar. 2022. "Study of Carrier Mobilities in 4H-SiC MOSFETS Using Hall Analysis" Materials 15, no. 19: 6736. https://doi.org/10.3390/ma15196736

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