Next Article in Journal
UAV Forensic Analysis and Software Tools Assessment: DJI Phantom 4 and Matrice 210 as Case Studies
Next Article in Special Issue
Temperature Dependent Analytical Model for the Threshold Voltage of the SiC VJFET with a Lateral Asymmetric Channel
Previous Article in Journal
Wideband Reconfigurable Integrated Low-Pass Filter for 5G Compatible Software Defined Radio Solutions
Previous Article in Special Issue
Role of Wide Bandgap Materials in Power Electronics for Smart Grids Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study and Assessment of Defect and Trap Effects on the Current Capabilities of a 4H-SiC-Based Power MOSFET

by
Fortunato Pezzimenti
1,*,
Hichem Bencherif
2,
Giuseppe De Martino
1,
Lakhdar Dehimi
2,
Riccardo Carotenuto
1,
Massimo Merenda
1 and
Francesco G. Della Corte
1
1
DIIES, Mediterranea University of Reggio Calabria, 89135 Reggio Calabria, Italy
2
LMSM, University of Biskra, Biskra 07000, Algeria
*
Author to whom correspondence should be addressed.
Electronics 2021, 10(6), 735; https://doi.org/10.3390/electronics10060735
Submission received: 22 February 2021 / Revised: 11 March 2021 / Accepted: 14 March 2021 / Published: 19 March 2021

Abstract

:
A numerical simulation study accounting for trap and defect effects on the current-voltage characteristics of a 4H-SiC-based power metal-oxide-semiconductor field effect transistor (MOSFET) is performed in a wide range of temperatures and bias conditions. In particular, the most penalizing native defects in the starting substrate (i.e., EH6/7 and Z1/2) as well as the fixed oxide trap concentration and the density of states (DoS) at the 4H-SiC/SiO2 interface are carefully taken into account. The temperature-dependent physics of the interface traps are considered in detail. Scattering phenomena related to the joint contribution of defects and traps shift the MOSFET threshold voltage, reduce the channel mobility, and penalize the device current capabilities. However, while the MOSFET on-state resistance (RON) tends to increase with scattering centers, the sensitivity of the drain current to the temperature decreases especially when the device is operating at a high gate voltage (VGS). Assuming the temperature ranges from 300 K to 573 K, RON is about 2.5 MΩ·µm2 for VGS > 16 V with a percentage variation ΔRON lower than 20%. The device is rated to perform a blocking voltage of 650 V.

1. Introduction

Silicon carbide (SiC) is worldwide recognized as a semiconductor well suited for high-temperature and high-power applications. In particular, the 4H-SiC polytype presents a high thermal conductivity on the order of 3–4 W/Kcm, a high specific resistivity of about 1011 Ω·cm, and a wide bandgap close to 3.23 eV at room temperature. Moreover, 4H-SiC-based devices are characterized by high critical electric fields and low leakage currents [1].
In modern power electronics, metal-oxide-semiconductor field effect transistors (MOSFETs) are widely valued for their low ON-state resistance (RON), high efficiency, and noticeable switching capabilities. Typical 4H-SiC-based MOSFETs are designed to support high blocking voltages ranging from 600 V to 1.7 kV [2,3,4]. However, the fundamental electrical parameters of a MOSFET, namely the breakdown voltage, output current, and specific RON, could be heavily affected by explicit trap/defect concentrations located in the bulk as well as in correspondence of the inversion layer at the silicon oxide (SiO2) interface [5,6,7,8].
From the literature, several papers have dealt with the 4H-SiC technological issues related, for example, to stacking faults, screw dislocations, and micro-pipes [9,10,11,12]. In more detail, in the MOSFET structure, a high density of states (DoS) at the 4H-SiC/SiO2 interface tends to prevent the realization of an efficient conductive channel, hence, reducing the carrier mobility. At the same time, the material intrinsic defects, in dependence of their capture cross sections, act as the primary carrier-lifetime killer in the drift region. Therefore, the presence of carbon atoms leads SiC-based devices to face higher concentrations of defects and traps in comparison with the conventional silicon technology. Obviously, these concentrations strongly depend on the effective quality of the starting materials and gate oxide interfaces.
The aim of this work is to assess the impact of trap and defect effects on the current-voltage characteristics of a power MOSFET in 4H-SiC. In particular, by means of a detailed numerical simulation study carried out at different temperatures (300 ≤ T ≤ 573 K) and bias conditions, the joint contribution of defects and traps is investigated accounting for their fundamental physical parameters such as the charge density, the location inside the bandgap, and the occupation probability. The physics of the interface trap distribution is modelled as temperature-dependent. This dependence results in a DoS spreading near the conduction and valence band edges for increasing values of T.
The device is dimensioned for a breakdown voltage (BVDS) of 650 V that meets the specifications of a huge market of power devices useful for several applications with special technical specifications (e.g., small size, thermal stability, low static power dissipation, ruggedness, 365-days-per-year operation under all weathers, etc.) which could be really satisfied by SiC technology. The MOSFET RON and its percentage variation with temperature (ΔRON) are considered key performance indicators during the simulations. In particular, RON is around 2.5 MΩ·µm2 with ΔRON in the limit of 20% for T ranging from 300 K to 573 K when imposing a gate voltage (VGS) higher than 16 V.
The presented study further extends the modelling efforts reported in recent authors’ manuscripts [13,14,15,16,17] where, by assuming the MOSFET structure trap/defect-free, we have explored the opportunity of down-shifting the SiC lower bound for a voltage rating around 150 V. There, in fact, while it has appeared rather evident that the use of SiC tends to lose in part its advantages for such extremely lower BVDS devices with respect to the use of silicon technology, preliminary results on the individual role of traps and defects have suggested the need to explore their combined effect in determining the effective MOSFET threshold voltage (VTH) and the RON behavior through the channel resistance and drift region contribution.

2. Device Structure

By using a 2D TCAD physical simulator [18], the MOSFET cross-section (half-cell) has been designed as shown in Figure 1. Although simplified for simulation purposes, this structure is in principle compatible with an actual 4H-SiC technological process based on doping by ion implantation [19,20,21].
Seven regions can be identified as follows. The drain is a heavily nitrogen-doped N+-region (region 1) that coincides with the 4H-SiC substrate on which the drift layer (region 2) is grown by epitaxy. Region 3 is the aluminum-doped p-base where the MOS structure and the conductive channel under the gate oxide lie. Region 4 is the phosphorous-doped source region. Region 5 is the insulating SiO2. Finally, region 6 and region 7 form the source and the gate contacts, respectively. Note that the source contact shorts the source and base regions to prevent the switch-on of the parasitic substrate(N+)-epilayer(N)-base(P)–source(N+) bipolar junction transistor.
Referring to Figure 1, the geometrical parameters and doping concentrations of the different MOSFET regions are summarized in Table 1. The half-cell width is 7.5 µm and the simulated device footprint is 7.5 µm2.
The distance between the p-base regions, Wj, is set to 5 µm (accumulation region) while the Wdrift thickness (5 μm) assures a MOSFET breakdown voltage close to 650 V as verified in [13]. This result is consistent with the calculations that we can perform by adapting to the P-base/N-epilayer/N+-substrate structure the standard expression valid for an abrupt junction p-i-n diode in punch-through condition [22], i.e.,
B V D S = E C W d r i f t q N e p i ( W d r i f t ) 2 2 ε s
where EC = 2 MV/cm is the 4H-SiC critical electric field (typical value), q is the electron charge, and εs is the semiconductor dielectric constant. From the theory, the lower the desired BVDS the higher the drift layer doping (Nepi), considering that, for a 4H-SiC-based MOSFET, Nepi is generally in the range 5 × 1015–1016 cm−3 [20,21,23].
For the proposed device, the overall ON-state resistance is determined by the sum of six different terms:
R O N = R N + + R c h + R a c c + R J F E T + R e p i + R s u b
where Rch is the channel resistance, RN+ is the source resistance, Racc is the accumulation layer resistance, RJFET is the resistance of the depletion layer between the P-base region and the N-epilayer, Repi is the epilayer region resistance, and Rsub is the substrate resistance. The contributions RN+ and Rsub are generally negligible because they are related to heavily doped regions. At the same time, Rch and Racc mainly depend on the gate bias level whereas RJFET and Repi are determined by the epilayer geometry and doping concentration.

3. Physical Models

The key physical models considered in this work include the standard expressions summarized in Table 2 [24,25,26,27,28,29,30,31,32].
The numerical simulation study solves the Poisson’s equation and the carrier continuity equations for a finely meshed device structure [18]. In particular, a mesh spacing down to 0.5 nm is imposed around the MOSFET p-n junctions and within the channel region. For the models in Table 2, all the reference parameters are reported in recent papers of ours focused on different 4H-SiC-based devices and supported by experimental results [14,33,34,35,36,37,38].
The carrier mobility degradation in the inversion layer, which is due to different scattering phenomena (e.g., Coulomb scattering and surface scattering effects), is modelled considering the transverse electric field component for carriers ( E ) by means of the expression [18]
μ n , p ( E ) = μ n , p ( 1 + E E n , p c r i t ) 1 2
where E n , p c r i t is an adjustable parameter.
In addition to the simulation setup recalled above, in the presented analysis the defect/trap effects are carefully taken into account as described in the following. In fact, the quality of the 4H-SiC/SiO2 interface as well as the presence of deep defect states in the 4H-SiC substrate and epilayer are crucial factors in determining the MOSFET electrical characteristics.

3.1. H-SiC/SiO2 Interface Traps

In a poor-quality 4H-SiC/SiO2 interface, a large number of interface-trapped charges related to traps having energy levels inside the semiconductor bandgap strongly affect the current capabilities of a physical device. The total amount of this trapped charges basically depends on the location of the intrinsic Fermi level, and an exchange of carriers with the valence and conduction bands takes place in turn through the trap capture and emission rates. In this study, the DoS induced by traps at the 4H-SiC/SiO2 interface is modelled as a sum of different contributions, namely a deep level distribution of states in the midgap (DMid), which presents a constant density with energy, and two exponentially decaying band tail states close to the conduction and valence band-edges (DTC, DTV) acting either as acceptor- or donor-like levels for free carriers [18]. As well known, acceptor-like centers are neutral when empty and become negatively charged when filled while donor-like centers are positively charged when empty and become neutral after capturing electrons (emitting holes). The model expressions for the tail state densities are in the form of
D T C ( E , T ) = D T C 0 exp ( E E C U C ( T ) )
D T V ( E , T ) = D T V 0 exp ( E V E U V ( T ) )
where UC and UV are the temperature-dependent characteristic energy decays of these profiles, and D T C 0 and D T V 0 are the tail state densities at the conduction band edge and valence band edge, respectively. This model is in accordance with different experimental results reported in the literature for different materials and dopants [39,40,41,42,43,44].
The DoS effects in the device channel region are accounted for by solving the following Poisson’s equation for the electrostatic potential (ψ)
( ε Ψ ) = q ( n p N D + + N A ) Q T
Here, in addition to the ionized impurity concentrations N D + and N A , which are expressed by the following standard expressions:
N D + = N D 1 + 2 exp ( E F n E D k T )
N A = N A 1 + 4 exp ( E A E F p k T )
where ND and NA are the substitutional n-type and p-type doping concentrations, ED and EA are the donor and acceptor energy levels, and EFn and EFp are the quasi-Fermi energy levels for electrons and holes, respectively, the trapped charge contribution QT is calculated as
Q T = q ( N t D + N t A )
considering that the ionized densities of traps N t D + (donor-like) and N t A (acceptor-like) are terms which depend on the product between the trap density and its probability of occupation, i.e.,
F t = v n , p σ n , p + e p , n v n , p ( σ p + σ n ) + ( e n + e p )
where vn,p is the carrier thermal velocity, σn,p is the carrier capture cross section, and en,p is the trap emission rate for electrons and holes given by
e n = v n σ n n i exp ( E E i k T )
e p = v p σ p n i exp ( E i E k T )
In Equations (11) and (12), ni is the material intrinsic carrier concentration and Ei is the intrinsic Fermi level.
The DoS parameters used in the simulations at T = 300 K are summarized in Table 3 [44,45,46]. According to experimental data from the literature, we can put in evidence that the band-edge trap densities of states are over two orders of magnitude higher than the energy-constant midgap contribution [42,43].
Finally, in the simulation setup we have considered a thin film of fixed oxide traps next to the 4H-SiC interface. Although the fixed oxide traps cannot exchange charge with carriers, due to their location they can act as Coulombic scattering centers that reduce the MOSFET threshold voltage and degrade the carrier mobility in the inversion layer. The effective concentration of the fixed oxide traps (Nfix) strictly depends on the device oxidation process and we have assumed typical values from the literature in the range 2 × 1011–1.3 × 1012 cm−2 [41,42,43].

3.2. H-SiC Intrinsic Defects

In the design of the MOSFET structure in Figure 1, the presence of 4H-SiC intrinsic defects cannot be neglected. In fact, the minority carrier current generated in the channel region flows vertically through the epilayer and the N+-substrate.
The most penalizing native defects in 4H-SiC-based devices are the so-called EH6/7 and Z1/2 centers [47]. These defects are both originated by a carbon vacancy due to, for example, electron radiations or high-temperature treatments. They are deep levels not independent each other and their concentration ratio is often considered unitary [48,49,50,51]. In more detail, Z1/2 centers are located in the upper half of the 4H-SiC bandgap whereas the EH6/7 level is close to the midgap (~1.65 eV). Although the EH6/7 nature as recombination centers can be uncertain, in the n-type 4H-SiC Z1/2 and EH6/7 are usually recognized as the acceptor-level and the donor-level of a carbon vacancy, respectively [45,47,48,49].
Table 4 summarizes the capture cross sections and the energy levels from the conduction band used during the simulations for the EH6/7 and Z1/2 centers. These parameters are consistent with several sets of data reported in the literature [48,51,52,53,54]. In particular, a wide range of defect concentrations is investigated starting from Nt = 2 × 1014 cm−3 up to Nt = 3 × 1015 cm−3.
It is worthwhile noting that to account for carrier lifetime killing effects caused by the 4H-SiC intrinsic defects, we have considered the carrier lifetimes governed both by the doping-dependent model reported in Table 2 and by the following expression [18]
τ n , p = 1 v n , p σ n , p N t
where the minority carrier lifetime is inversely correlated with Nt. This behavior has been verified experimentally for 4H-SiC epitaxial layers of growing thickness with defect concentrations exceeding 1 × 1013 cm−3 by deep level transient spectroscopy (DLTS) and minority carrier transient spectroscopy (MCTS) measurements [54,55].

4. Results and Discussion

The simulation analysis starts considering the 4H-SiC MOSFET described in Table 1 as a device free of intrinsic defects and interface traps (fresh device). The drain current (IDS)-drain voltage (VDS) output characteristics nearby and within the triode region of this device for 8 ≤ VGS ≤ 20 V are shown in Figure 2 (T = 300 K) and Figure 3 (T = 573 K). The value VGS = 8 V can be roughly assumed close to the effective MOSFET threshold voltage.
We can note that the drain current decreases harshly when increasing the temperature. This effect is mainly related to the temperature dependence of the carrier mobility in the inversion layer and drift region, determining an overall increase of the device ON-state resistance.
The calculated behaviors of RON as a function of VGS are shown in Figure 4. Here, the percentage variation with temperature (ΔRON) is also reported. In particular, ΔRON quickly tends to stabilize around 300%. As expected, an increasing value of T determines a change in the charge density both in the inversion and in the depletion region with a severe impact on the RON curve.
It is important to note that the prediction capabilities of the adopted simulation setup have been tested in [13] by comparing the RON results for an almost similar device (BVDS = 900 V) with the datasheet values of a commercial MOSFET in 4H-SiC [56]. In particular, a good agreement has been achieved calculating RON close to 520 kΩ×µm2 for VGS = 15 V and VDS = 1 V at room temperature.
Introducing different Z1/2 and EH6/7 intrinsic defect concentrations for 4H-SiC in the proposed MOSFET structure, we have calculated the IDS-VGS characteristics showed in Figure 5.
As stated previously, the output current mainly flows vertically through the epilayer and substrate and therefore it is evident that as Nt increases IDS decreases unavoidably. In particular, as verified during the simulations, Nt = 2 × 1014 cm−3 can be considered the limit value to preserve the drain current behavior in the whole explored VGS range. In fact, the more Nt tends to become comparable to the epilayer doping concentration (5 × 1015 cm−3) the more the device current capabilities are strongly penalized.
The main reason of the results in Figure 5 lies in the penalized flow of electrons due to the defect effects which originate in the MOSFET drift region increasing the local recombination rate. In other words, the drain current is degraded because these defects reduce the carrier lifetime and act as carrier traps in the device active region introducing high-resistive paths. The weight of the term Nt in penalizing the device current-voltage characteristics is almost the same for different temperatures when the MOSFET is firmly in ON-state. For example, in the 300–573 K temperature range, with respect to the fresh device, we can calculate always an IDS reduction on the order of 5% and 20% for Nt = 2 × 1014 cm−3 and Nt = 9 × 1014 cm−3, respectively.
On the contrary, the presence of 4H-SiC intrinsic defects affects the MOSFET threshold voltage variation with temperature differently. In fact, as shown in Figure 6, while VTH tends to increase with increasing Nt for a fixed temperature, this effect tends to become less prominent at the higher values of T (>400 K). In Figure 6, VTH is always calculated by imposing a subthreshold drain current in the limit of 10 nA.
The increase of the intrinsic defect concentration in the starting epilayer contributes to prevent the creation of the MOSFET conductive channel. In fact, defect effects enhance the recombination rate in the inversion layer via reducing the carrier lifetimes and excluding electrons from transport mechanisms. At each temperature, the filled traps originate Coulombic scattering phenomena that determine a positive shift of VTH and reduce the device output current. On the other hand, the more the temperature increases the more the number of filled traps decreases. Thus, the threshold voltage variation with Nt decreases.
The carrier mobility is another fundamental physical parameter strongly affected by temperature and defects. In more detail, the total carrier mobility is a sum of different contributions depending on the material doping concentration as well as on the local electric field in the device structure and the scattering mechanisms from interface charges and ionized impurities in the bulk. For the proposed MOSFET, the carrier mobility degradation in the inversion layer due to the intrinsic defect concentration at different temperatures is shown in Figure 7. With respect to the fresh device (Nt = 0), we can state that the increased recombination effects relate to the progressive increase of Nt have a severe impact on the channel mobility (µch) at any value of T.
The RON curves as a function of Nt for three different temperatures are shown in Figure 8.
The presence of 4H-SiC intrinsic defects increases RON considerably with an almost similar behavior in dependence of T. This result is more noticeable the closer Nt approaches (overcomes) the epilayer doping concentration (5 × 1015 cm−3). In fact, these defects act as efficient recombination centers both in the channel region and in the drift region causing higher resistive paths for the current flow. In particular, in a physical device, Z1/2 centers with their negative-U nature (i.e., the Z1 and Z2 energy levels are very close to each other) can be considered able to capture couples of electrons almost simultaneously. At the same time, the positively charged EH6/7 defects (when empty), with their larger cross sections for electrons, are deeply involved in the detrimental effects in the channel region of an n-type MOSFET.
After fixing Nt = 2 × 1014 cm−3, in order to refer the MOSFET analysis to a more realistic device, in the successive simulations the trap effects at the 4H-SiC/SiO2 interface have been accounted for. As expected, these interface traps heavily affect the device performance by degrading the channel mobility and increasing the channel resistance contribution further. In a first step, we have investigated the role of the fixed oxide traps by assuming, in accordance with experimental data [41,42,43], a reference value Nfix = 1 × 1012 cm−2 into a thin film of SiO2 located at the 4H-SiC interface. These centers, which number can be assumed independent from T, are positive charges responsible for the semiconductor band-bending at the interface [57,58]. The band-bending effect sustains the formation of the conductive channel for lower values of VGS and therefore tends to reduce the MOSFET threshold voltage as shown in Figure 9.
By comparing the IDS curves for Nfix = 0 and Nfix = 1 × 1012 cm−2 we can note a significant shift to lower gate voltages. These behaviors aid to clarify the key role of the fixed oxide traps in determining the effective MOSFET VTH both at low and high working temperature.
In a second step, we have finally involved in the simulations an explicit interface-trapped charge. In more detail, by assuming the DoS parameters listed in Table 3, from the simulations it was rather evident that the traps mainly affecting the MOSFET ON-state current capabilities are those located in the upper half of the bandgap. In fact, since the proposed device is a n-channel MOSFET operating in the inversion regime, the more the Fermi level moves in the upper half of the bandgap the more the traps with energetic states close to the conduction band govern the device electrical characteristics. On the other hand, in the subthreshold regime, when the Fermi level is close to the middle of the bandgap, it is the deep level distribution of states in the midgap to be occupied.
The tail traps effect on the MOSFET channel mobility and threshold voltage at T = 300 K is shown in Figure 10. In particular, starting in (4) from the reference value D0TC = 5.72 × 1013 cm−2 eV−1 [43], we have investigated a wide range of trap densities (1 × 1012–1 × 1014 cm−2 eV−1). It is important to note that the results in Figure 10 also involve the combined effects due to the fixed oxide charge (Nfix = 1 × 1012 cm−2) and the 4H-SiC intrinsic defect concentration (Nt = 2 × 1014 cm−3).
When increasing the tail traps, they are able to trap an increasing number of electrons in the near-threshold regime of the device, preventing the complete channel formation and thus determining a positive shift of VTH similarly to the intrinsic defect concentration effect. At the same time, the increased number of filled tail traps increasingly enhances the scattering phenomena of mobile charges in the inversion layer for VGS > VTH. For example, we can expect a channel mobility which is about 20 cm2/Vs for VGS = 12 V and VDS = 1 V. In particular, a tail trap density up to 1 × 1013 cm−2 eV−1 leads to a faster variation of the µch and VTH curves. From the literature, a D0TC value close to 1 × 1013 cm−2 eV−1 was extracted experimentally in [41]. However, other experimental works dealt with a tail trap density of about 6 × 1013 cm−2 eV−1 [42,43]. Obviously, the exact value is strictly dependent on the available technology and device fabrication process.
In principle, we can assume that the carrier mobility behavior in the MOSFET structure is controlled by Coulomb scattering for low gate voltages around VTH whereas for an increasing VGS, which increases the transverse electric field and more and more confines electrons at the 4H-SiC/SiO2 interface, the surface scattering effects in the channel region become dominant. In other words, for high gate voltages, due to an increased screening of the interface and fixed oxide traps by the inversion layer, Coulomb scattering is less effective.
By performing a detailed high-temperature analysis, interesting information on the interface trap effects can be extracted. In fact, although with an increasing temperature the fixed oxide charge can be assumed constant and the role of the midgap density of states remains negligible, the distribution of the band-edge interface trap densities is temperature-dependent through the band-tail energy parameter (U). In particular, while the temperature increases and the bandgap narrowing effect starts, the tail state profiles spread deeper into the bandgap with a reduction of the band-edge intercept densities. In more detail, to evaluate the MOSFET current capabilities at T = 573 K, we have used in (4) UC = 120 meV (67 meV @T = 300 K) and D0TC = 3.2 × 1013 cm−2eV−1 (5.7 × 1013 cm−2eV−1 @T = 300 K). These values are expected from the experimental data reported in [43] since their temperature dependence is almost monotonic.
Figure 11 shows the MOSFET RON behaviors calculated at low and high temperatures, assuming the joint contribution of all the defect/trap effects described previously. Here, the curves presented in Figure 4 for a fresh device are also reported for comparison.
It is worthwhile noting that, when the temperature increases from 300 K to 573 K, in contrast with the electrical characteristics of a fresh device, for a more realistic MOSFET the difference in the RON curves tends to reduce with VGS. In particular, at the higher gate biases RON increasingly becomes slightly dependent on temperature although a large variation of T is considered. For example, RON ≈ 2.5 MΩ·µm2 with a ΔRON close to 10% for VGS = 18 V. This limited RON increase with temperature for a MOSFET with defects and traps represents a key finding of the presented analysis. It can be explained by considering that while Coulomb scattering is directly proportional to the occupied density of states, it is inversely proportional to the device operation temperature. In other words, since mobile carriers have more energy with increasing values of T, they are less affected by the Coulomb potential at the 4H-SiC interface and, therefore, for a given ON-state bias condition, much more electrons are available for conduction in the inversion layer. From the simulations, the more the number of the interface trap states is large the more this effect is evident in counteracting the current reduction due to the lower carrier mobility in the channel region as verified for the fresh device.
In sum, the observed results in Figure 11 depend on both the occupied density of states and temperature. In particular, the fresh device presents a RON curve with a positive temperature coefficient that is almost constant in the whole explored VGS range (i.e., ΔRON ≈ 300% @T = 573 K) whereas a MOSFET structure involving intrinsic defects and an explicit DoS at the gate oxide interface presents a decreasing ΔRON in dependence of VGS. Obviously, the presence of defects and traps leads the MOSFET to operate with a lower drain current at any fixed bias voltage and temperature. As an additional result note that, with respect to the curves in Figure 11, by reducing the trapped charge density by a factor of 10 we have calculated an RON improvement on the order of 15% preserving about the same ΔRON. This improvement is certainly due to an increased inversion charge and an enhanced channel mobility.
Finally, it could be interesting to cite the state-of-the-art of Si-based super-junction power devices rated in the 600–700 V class which, while showing a specific RON around 2 MΩ·µm2 at room temperature [59], suffer for a severe performance degradation with TRON ≈ 150% @T = 423 K) under test conditions [60]. On the other hand, datasheet values in terms of ΔRON for similar 4H-SiC-based MOSFETs [2,3,4] confirm the results obtained in this paper.

5. Conclusions

An exhaustive simulation study on the electrical characteristics of a power MOSFET in 4H-SiC is presented, accounting for trap and defect effects which originate in the bulk and at the gate oxide interface. Starting from a fresh device, the joint contribution of defects and traps determines a significative shift of the threshold voltage and increases the MOSFET RON at any working temperature. In more detail, the intrinsic defect concentration should result at least one order of magnitude lower than the epilayer doping concentration to avoid the formation of high-resistive paths for current. At the same time, the temperature-dependent physics of the density of states located at the SiO2 interface plays a key role in determining the MOSFET output characteristics. In fact, in contrast with the electrical behavior of a fresh device, in presence of defects and traps, the calculated increase of RON with temperature decreases for increasing values of VGS. In particular, for VGS higher than 16 V, the RON value remains around 2.5 MΩ·µm2 with a percentage variation in the limit to 20% although a large excursion of temperature is considered (300–573 K). The presented analysis turns useful to assess the effects of different scattering phenomena in the device inversion layer and drift region, thus supporting a clear understanding of the proven reliability of 4H-SiC-based MOSFETs for high-temperature applications.

Author Contributions

Conceptualization, F.P., L.D. and F.G.D.C.; methodology and validation, F.P., H.B., L.D. and R.C.; software, H.B., G.D.M. and M.M.; investigation, F.P. and H.B.; data curation, F.P., G.D.M. and M.M.; visualization, F.P. and R.C.; writing—original draft preparation, F.P.; writing—review and editing, F.P.; formal analysis and supervision, F.P., L.D. and F.G.D.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

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. SiC MOSFETs. Available online: https://www.rohm.com/products/sic-power-devices/sic-mosfet.
  3. Discrete SiC MOSFETs. Available online: https://www.wolfspeed.com/power/products/sic-mosfets.
  4. Silicon Carbide CoolSiC™ MOSFETs. Available online: https://www.infineon.com/cms/en/product/power/mosfet/silicon-carbide/.
  5. Tachiki, H.; Ono, T.; Kobayashi, T.; Tanaka, H.; Kimoto, T. Estimation of threshold voltage in SiC short-channel MOSFETs. IEEE Trans. Electron. Dev. 2018, 65, 3077–3080. [Google Scholar] [CrossRef]
  6. Tanimoto, Y.; Saito, A.; Matsuura, K.; Kikuchihara, H.; Mattausch, H.J.; Miura-Mattausch, M.; Kawamoto, N. Power-loss prediction of high-voltage SiC-mosfet circuits with compact model including carrier-trap influences. IEEE Trans. Power Electron. 2016, 31, 4509–4516. [Google Scholar] [CrossRef]
  7. Ettisserry, D.P.; Goldsman, N.; Lelis, A. A methodology to identify and quantify mobility-reducing defects in 4H-silicon carbide power metal-oxide-semiconductor field-effect transistors. J. Appl. Phys. 2014, 115, 103706. [Google Scholar] [CrossRef]
  8. Haasmann, D.; Dimitrijev, S. Energy position of the active near-interface traps in metal-oxide-semiconductor field-effect transistors on 4H-Sic. Appl. Phys. Lett. 2013, 103, 1135061–1135063. [Google Scholar] [CrossRef] [Green Version]
  9. Hornos, T.; Gali, A.; Svensson, B.G. Negative-U system of carbon vacancy in 4H-SiC. Mater. Sci. Forum 2011, 679, 261–264. [Google Scholar] [CrossRef] [Green Version]
  10. Danno, K.; Nakamura, D.; Kimoto, T. Investigation of carrier lifetime in 4H-SiC epilayers and lifetime control by electron irradiation. Appl. Phys. Lett. 2007, 90, 202109. [Google Scholar] [CrossRef] [Green Version]
  11. Kawahara, K.; Trinh, X.T.; Son, N.T.; Janzen, E.; Suda, J.; Kimoto, T. Quantitative comparison between Z1/2 center and carbon vacancy in 4H-SiC. J. Appl. Phys. 2014, 115, 143705. [Google Scholar] [CrossRef] [Green Version]
  12. Booker, I.D.; Janzén, E.; Son, N.T.; Hassan, J.; Stenberg, P.; Sveinbjörnsson, E.Ö. Donor and double-donor transitions of the carbon vacancy related EH6∕7 deep level in 4H-SiC. J. Appl. Phys. 2016, 119, 235703. [Google Scholar] [CrossRef]
  13. Della Corte, F.G.; De Martino, G.; Pezzimenti, F.; Adinolfi, G.; Graditi, G. Numerical simulation study of a low breakdown voltage 4H-SiC MOSFET for photovoltaic module-level applications. IEEE Trans. Electron. Devices 2018, 65, 3352–3360. [Google Scholar] [CrossRef]
  14. Bencherif, H.; Dehimi, L.; Pezzimenti, F.; De Martino, G.; Della Corte, F.G. Multiobjective optimization of design of 4H-SiC power MOSFETs for specific applications. J. Electron. Mater. 2019, 48, 3871–3880. [Google Scholar] [CrossRef]
  15. De Martino, G.; Pezzimenti, F.; Della Corte, F.G. Interface trap effects in the design of a 4H-SiC MOSFET for low voltage applications. In Proceedings of the 2018 International Semiconductor Conference (CAS), Sinaia, Romania, 10–12 October 2018. [Google Scholar]
  16. Bencherif, H.; Pezzimenti, F.; Dehimi, L.; Della Corte, F. Analysis of 4H-SiC MOSFET with distinct high-k/4H-SiC interfaces under high temperature and carrier-trapping conditions. Appl. Phys. A-Mater. 2020, 126, 854. [Google Scholar] [CrossRef]
  17. Bencherif, H.; Dehimi, L.; Athamena, N.E.; Pezzimenti, F.; Megherbi, M.L.; Della Corte, F.G. Simulation Study of Carbon Vacancy Trapping Effect on Low Power 4H-SiC MOSFET Performance. Silicon 2021, 1–9. [Google Scholar] [CrossRef]
  18. Silvaco Int. Atlas User’s Manual, Device Simulator Software. Available online: https://www.eng.buffalo.edu/~wie/silvaco/atlas_user_manual.pdf (accessed on 2 October 2004).
  19. Okamoto, M.; Iijima, M.; Nagano, T.; Fukuda, K.; Okumura, H. Controlling characteristics of 4H-SiC (0001) p-channel MOSFETs fabricated on ion-implanted n-well. Mater. Sci. Forum 2012, 717, 781–784. [Google Scholar] [CrossRef]
  20. Sung, W.; Baliga, B.J. Monolithically integrated 4H-SiC MOSFET and JBS diode (JBSFET) using a single Ohmic/Schottky process scheme. IEEE Electron. Device Lett. 2016, 37, 1605–1608. [Google Scholar] [CrossRef]
  21. Mikamura, Y.; Hiratsuka, K.; Tsuno, T.; Michikoshi, H.; Tanaka, S.; Masuda, T.; Sekiguchi, T. Novel designed SiC devices for high power and high efficiency systems. IEEE Trans. Electron. Devices 2014, 62, 382–389. [Google Scholar] [CrossRef]
  22. Baliga, B.J. Fundamentals of Power Semiconductor Devices; Springer: New York, NY, USA, 2008. [Google Scholar]
  23. Harada, S.; Okamoto, M.; Yatsuo, T.; Adachi, K.; Fukuda, K.; Arai, K. 8.5-mΩ·cm2 600-V Double-Epitaxial MOSFETs in 4H–SiC. IEEE Electron. Device Lett. 2004, 25, 292–294. [Google Scholar] [CrossRef]
  24. Selberherr, S. Analysis and Simulation of Semiconductor Devices; Springer: New York, NY, USA, 1984. [Google Scholar]
  25. Li, X.; Luo, Y.; Fursin, L.; Zhao, J.H.; Pan, M.; Alexandrov, P.; Weiner, M. On the temperature coefficient of 4H-SiC BJT current gain. Solid State Electron. 2003, 47, 233–239. [Google Scholar] [CrossRef]
  26. Ruff, M.; Mitlehner, H.; Helbig, R. SiC devices physics and numerical simulation. IEEE Trans. Electron. Devices 1994, 41, 1040–1054. [Google Scholar] [CrossRef]
  27. Lindefelt, U. Doping-induced band edge displacements and band gap narrowing in 3C-, 4H-, 6H-SiC, and Si. J. Appl. Phys. 1998, 84, 2628–2637. [Google Scholar] [CrossRef]
  28. Landsberg, P.T.; Kousik, G.S. The connection between carrier lifetime and doping density in nondegenerate semiconductors. J. Appl. Phys. 1984, 56, 1696–1700. [Google Scholar] [CrossRef]
  29. Roschke, M.; Schwierz, F. Electron mobility models for 4H, 6H, and 3C SiC. IEEE Trans. Electron. Devices 2001, 48, 1442–1447. [Google Scholar] [CrossRef]
  30. Raghunathan, R.; Baliga, B.J. Measurement of electron and hole impact ionization coefficients for SiC. In Proceedings of the 9th International Symposium on Power Semiconductor Devices and IC’s, Weimar, Germany, 26–29 May 1997. [Google Scholar]
  31. Pezzimenti, F.; Bencherif, H.; Yousfi, A.; Dehimi, L. Current-voltage analytical model and multiobjective optimization of design of a short channel gate-all-around-junctionless MOSFET. Solid State Electron. 2019, 161, 107642. [Google Scholar] [CrossRef]
  32. Galeckas, A.; Linnros, J.; Grivickas, V.; Lindefelf, U.; Hallin, C. Auger recombination in 4H-SiC: Unusual temperature behaviour. Appl. Phys. Lett. 1997, 71, 3269–3271. [Google Scholar] [CrossRef]
  33. Zeghdar, K.; Bencherif, H.; Dehimi, L.; Pezzimenti, F.; Della Corte, F.G. Simulation and analysis of the forward bias current–voltage–temperature characteristics of W/4H-SiC Schottky barrier diodes for temperature-sensing applications. Solid State Electron. Lett. 2020, 2, 49–54. [Google Scholar] [CrossRef]
  34. Pezzimenti, F. Modeling of the steady state and switching characteristics of a normally-off 4H-SiC trench bipolar-mode FET. IEEE Trans. Electron. Devices. 2013, 60, 1404–1411. [Google Scholar] [CrossRef]
  35. Megherbi, M.L.; Pezzimenti, F.; Dehimi, L.; Saadoune, A.; Della Corte, F.G. Analysis of the forward I-V characteristics of Al-implanted 4H-SiC p-i-n diodes with modeling of recombination and trapping effects due to intrinsic and doping-induced defect states. J. Electron. Mater. 2018, 47, 1414–1420. [Google Scholar] [CrossRef]
  36. Pezzimenti, F.; Della Corte, F.G. Design and modeling of a novel 4H-SiC normally-off BMFET transistor for power applications. In Proceedings of the Melecon 2010–2010 15th IEEE Mediterranean Electrotechnical Conference, Valletta, Malta, 26–28 April 2010. [Google Scholar]
  37. Megherbi, M.L.; Pezzimenti, F.; Dehimi, L.; Rao, S.; Della Corte, F.G. Analysis of different forward current—Voltage behaviours of Al implanted 4H-SiC vertical p–i–n diodes. Solid State Electron. 2015, 109, 12–16. [Google Scholar] [CrossRef]
  38. Zeghdar, K.; Dehimi, L.; Pezzimenti, F.; Megherbi, M.L.; Della Corte, F.G. Analysis of the electrical characteristics of Mo/4H-SiC Schottky barrier diodes for temperature-sensing applications. J. Electron. Mater. 2020, 49, 1322–1329. [Google Scholar] [CrossRef]
  39. Sharma, P.; Minakshi, M.; Watcharatharapong, T.; Laird, D.; Euchner, H.; Ahuja, R. Zn Metal Atom Doping on the Surface Plane of One-Dimesional NiMoO4 Nanorods with Improved Redox Chemistry. ACS Appl. Mater. Interfaces 2020, 12, 44815–44829. [Google Scholar] [CrossRef]
  40. Sharma, P.; Minakshi, M.; Singh, D.; Ahuja, R. Highly Energetic and Stable Gadolinium/Bismuth Molybdate with a Fast Reactive Species, Redox Mechanism of Aqueous Electrolyte. ACS Appl. Energy Mater. 2020, 3, 12385–12399. [Google Scholar] [CrossRef]
  41. Dhar, S.; Haney, S.; Cheng, L.; Ryu, S.R.; Agarwal, A.K. Inversion layer carrier concentration and mobility in 4H-SiC metal-oxide-semiconductor field-effect transistors. J. Appl. Phys. 2010, 108, 054509. [Google Scholar] [CrossRef]
  42. Potbhare, S.; Goldsman, N.; Pennington, G. Numerical and experimental characterization of 4H-silicon carbide lateral metal-oxide-semiconductor field-effect transistors. J. Appl. Phys. 2006, 100, 044515. [Google Scholar] [CrossRef]
  43. Potbhare, S.; Goldsman, N.; Lelis, A.; McGarrity, J.M.; McLean, F.B. A physical model of high temperature 4H-SiC MOSFETs. IEEE Trans. Electron. Devices 2008, 55, 2029–2039. [Google Scholar] [CrossRef]
  44. Afanasev, V.; Bassler, M.; Pensl, G.; Schulz, M. Intrinsic SiC/SiO2 interface states. Phys. Status Solidi A 1997, 162, 321–337. [Google Scholar] [CrossRef]
  45. Dimitriadis, E.I.; Archontas, N.; Girginoudi, D.; Georgoulas, N. Two dimensional simulation and modeling of the electrical characteristics of the a-SiC/c-Si(p) based, thyristor-like, switches. Microelectron. Eng. 2015, 133, 120–128. [Google Scholar] [CrossRef]
  46. Goldberg, Y.; Levinshtein, M.E.; Rumyantsev, S.L. Properties of Advanced Semiconductor Materials GaN, AlN, SiC, BN, SiC, SiGe; J. Wiley & Sons: New York, NY, USA, 2001. [Google Scholar]
  47. Feng, Z.C.; Zhao, J.H. Silicon Carbide: Materials, Processing and Devices; Taylor & Francis: New York, NY, USA, 2004. [Google Scholar]
  48. Son, N.T.; Trinh, X.T.; Løvlie, L.S.; Svensson, B.G.; Kawahara, K.; Suda, J.; Kimoto, T.; Umeda, T.; Isoya, J.; Makino, T.; et al. Negative-U System of Carbon Vacancy in 4H-SiC. Phys. Rev. Lett. 2012, 109, 187603. [Google Scholar] [CrossRef] [Green Version]
  49. Danno, K.; Kimoto, T. Investigation of deep levels in n-type 4H-SiC epilayers irradiated with low-energy electrons. J. Appl. Phys. 2006, 100, 113728. [Google Scholar] [CrossRef]
  50. Zippelius, B.; Glas, A.; Weber, H.B.; Pensl, G.; Kimoto, T.; Krieger, M. Z1/2- and EH6-center in 4H-SiC: Not identical defects. Mater. Sci. Forum 2014, 717, 251–254. [Google Scholar]
  51. Hornos, T.; Gali, A.; Svensson, B.G. Large-scale electronic structure calculations of vacancies in 4H-SiC using the Heyd-Scuseria-Ernzerhof screened hybrid density functional. Mater. Sci. Forum 2011, 679, 261–264. [Google Scholar] [CrossRef] [Green Version]
  52. Mitchel, W.C.; Mitchell, W.D. Compensation mechanism in high purity semi-insulating 4H-SiC. J. Appl. Phys. 2007, 101, 053716. [Google Scholar] [CrossRef] [Green Version]
  53. Klein, P.B. Carrier lifetime measurements in n- 4H-SiC epilayers. J. Appl. Phys. 2008, 103, 033702. [Google Scholar] [CrossRef]
  54. Klein, P.B.; Shanabrook, B.V.; Huh, S.W.; Polyakov, A.Y.; Skowronski, M.; Sumakeris, J.J.; O’Loughlin, M.J. Lifetime-limiting defects in n- 4H-SiC epilayers. Appl. Phys. Lett. 2006, 88, 052110. [Google Scholar] [CrossRef] [Green Version]
  55. Booker, I.D.; Okuda, T.; Grivickas, P.; Hassan, J.; Janzén, E.; Sveinbjörnsson, E.Ö.; Suda, J.; Kimoto, T. Device-relevant and processing induced traps and recombination centers in 4H-SiC. In Proceedings of the 11th European Conference on Silicon Carbide and Related Materials 2016, Halkidiki, Greece, 25–29 September 2016. [Google Scholar]
  56. CREE Model C3M0280090D. Available online: http://www.cree.com.
  57. Kerber, A.; Cartier, E.; Pantisano, L.; Degraeve, R.; Kauerauf, T.; Kim, Y.; Hou, A.; Groeseneken, G.; Maes, H.E.; Schwalke, U. Origin of the threshold voltage instability in SiO2/HfO2 dual layer gate dielectrics. IEEE Electr. Device Lett. 2003, 24, 87–89. [Google Scholar] [CrossRef]
  58. Zafar, S.; Callegari, A.; Gusev, E.; Fischetti, M.V. Charge trapping related threshold voltage instabilities in high permittivity gate dielectric stacks. J. Appl. Phys. 2003, 93, 9298. [Google Scholar] [CrossRef]
  59. Udrea, F.; Deboy, G.; Fujihira, T. Superjunction power devices, history, development, and future prospects. IEEE Trans. Electron. Devices 2017, 64, 713–727. [Google Scholar] [CrossRef]
  60. 600 V and 650 V CoolMOS™ C6/E6. Available online: https://www.infineon.com/cms/en/product/power/mosfet/500v-900v-coolmos-n-channel-power-mosfet/600v-and-650v-coolmos-c6-e6/.
Figure 1. Schematic cross-sectional view of the metal-oxide-semiconductor field effect transistor (MOSFET) half-cell. The drawing is not to scale.
Figure 1. Schematic cross-sectional view of the metal-oxide-semiconductor field effect transistor (MOSFET) half-cell. The drawing is not to scale.
Electronics 10 00735 g001
Figure 2. Forward IDS-VDS characteristics of the device in Table 1 at T = 300 K.
Figure 2. Forward IDS-VDS characteristics of the device in Table 1 at T = 300 K.
Electronics 10 00735 g002
Figure 3. Forward IDS-VDS characteristics of the device in Table 1 at T = 573 K.
Figure 3. Forward IDS-VDS characteristics of the device in Table 1 at T = 573 K.
Electronics 10 00735 g003
Figure 4. ON-state resistance versus VGS for a fresh device. VDS = 1 V.
Figure 4. ON-state resistance versus VGS for a fresh device. VDS = 1 V.
Electronics 10 00735 g004
Figure 5. IDS-VGS curves for different 4H-SiC intrinsic defect concentrations. VDS = 5 V.
Figure 5. IDS-VGS curves for different 4H-SiC intrinsic defect concentrations. VDS = 5 V.
Electronics 10 00735 g005
Figure 6. Threshold voltage behavior as a function of Nt at different temperatures.
Figure 6. Threshold voltage behavior as a function of Nt at different temperatures.
Electronics 10 00735 g006
Figure 7. Channel mobility as a function of Nt at different temperatures. VGS = 12 V, VDS = 1 V.
Figure 7. Channel mobility as a function of Nt at different temperatures. VGS = 12 V, VDS = 1 V.
Electronics 10 00735 g007
Figure 8. ON-state resistance as a function of Nt at different temperatures. VGS = 12 V, VDS = 5 V.
Figure 8. ON-state resistance as a function of Nt at different temperatures. VGS = 12 V, VDS = 5 V.
Electronics 10 00735 g008
Figure 9. IDS-VGS curves for different values of Nfix. VDS = 1V.
Figure 9. IDS-VGS curves for different values of Nfix. VDS = 1V.
Electronics 10 00735 g009
Figure 10. Channel mobility (VGS = 12 V, VDS = 1 V) and device threshold voltage as a function of the tail trap density. DMid = 2.4 × 1011 cm−2 eV−1.
Figure 10. Channel mobility (VGS = 12 V, VDS = 1 V) and device threshold voltage as a function of the tail trap density. DMid = 2.4 × 1011 cm−2 eV−1.
Electronics 10 00735 g010
Figure 11. ON-state resistance versus VGS for alternative devices with and without defect/trap effects. VDS = 1 V.
Figure 11. ON-state resistance versus VGS for alternative devices with and without defect/trap effects. VDS = 1 V.
Electronics 10 00735 g011
Table 1. MOSFET structure.
Table 1. MOSFET structure.
ParameterValue
Oxide thickness, tox (μm)0.08
Channel length, Lch (μm)1
Base junction depth, Wbase (μm)1.5
Interspace, W’base (μm)1
Base-to-base distance, Wj (μm)5
Epilayer thickness, Wdrift (μm)5
Base-to-substrate distance, W’drift (μm)3.5
Substrate thickness, Wsub (μm)100
N+-source doping (cm−3)1 × 1018
P-base doping (cm−3)1 × 1017
N-epilayer doping (cm−3)5 × 1015
N+-substrate doping (cm−3)1 × 1019
Table 2. Physical models for numerical simulation.
Table 2. Physical models for numerical simulation.
ModelExpression
Bandgap energy E g ( T ) = E g 300 θ ( T 300 )
Auger recombination R A u g e r = ( C A p p + C A n n ) ( n p n i 2 )
Shockley Read Hall recombination R S R H = p n n i 2 τ p ( n + n i exp ( E t r a p k T ) ) + τ n ( p + n i exp ( E t r a p k T ) )
Incomplete ionization of impurities N A , D + = N A , D ( 1 + 1 + 4 g v , c N A , D N V , C ( T ) e Δ E A , D k T 2 g v , c N A , D N V , C ( T ) e Δ E A , D k T )
Apparent bandgap narrowing Δ E g p , n = A p , n ( N A , D + 10 18 ) 1 2 + B p , n ( N A , D + 10 18 ) 1 3 + C p , n ( N A , D + 10 18 ) 1 4
Impact ionization Γ n , p = a 0 n , p exp ( b 0 n , p E )
Carrier lifetime τ n , p = τ 0 n , p ( T 300 ) λ 1 + ( N N n , p S R H )
Carrier mobility for low-field and
high-field conditions
μ n , p = μ 0 n , p min ( T 300 ) α + μ 0 n , p max ( T 300 ) β μ 0 n , p min ( T 300 ) α 1 + ( T 300 ) γ ( N N n , p c r i t ) δ μ n , p ( E ) = μ n , p [ 1 + ( E μ n , p v s a t ( T ) ) κ ] 1 κ
Saturation velocity v s a t ( T ) = v s a t 300 + ( 300 T ) η
Table 3. DoS reference parameters at T = 300 K.
Table 3. DoS reference parameters at T = 300 K.
ParameterValue
Tail state density, D0TC,V (cm−2 eV−1)5.7 × 1013
Band-tail energy decay, UC,V (meV)67
Mid-gap density, DMid (cm−2 eV−1)2.4 × 1011
Thermal velocity, vn,p (cm/s)1.9 × 107, 1.2 × 107
Capture cross section, σnA,D (cm2)1 × 10−16, 1 × 10−14
Capture cross section, σpA,D (cm2)1 × 10−14, 1 × 10−16
Table 4. Intrinsic defect parameters.
Table 4. Intrinsic defect parameters.
Et (eV)Nt (cm−3)σn (cm2)σp (cm2)
Z1/20.672 × 1014–3 × 10152.0 × 10−143.5 × 10−14
EH6/71.652 × 1014–3 × 10152.4 × 10−141.0 × 10−14
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pezzimenti, F.; Bencherif, H.; De Martino, G.; Dehimi, L.; Carotenuto, R.; Merenda, M.; Della Corte, F.G. Study and Assessment of Defect and Trap Effects on the Current Capabilities of a 4H-SiC-Based Power MOSFET. Electronics 2021, 10, 735. https://doi.org/10.3390/electronics10060735

AMA Style

Pezzimenti F, Bencherif H, De Martino G, Dehimi L, Carotenuto R, Merenda M, Della Corte FG. Study and Assessment of Defect and Trap Effects on the Current Capabilities of a 4H-SiC-Based Power MOSFET. Electronics. 2021; 10(6):735. https://doi.org/10.3390/electronics10060735

Chicago/Turabian Style

Pezzimenti, Fortunato, Hichem Bencherif, Giuseppe De Martino, Lakhdar Dehimi, Riccardo Carotenuto, Massimo Merenda, and Francesco G. Della Corte. 2021. "Study and Assessment of Defect and Trap Effects on the Current Capabilities of a 4H-SiC-Based Power MOSFET" Electronics 10, no. 6: 735. https://doi.org/10.3390/electronics10060735

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