c08f023

Figure 8.23 Doping profile and band diagram for MOSFETs with an n-type surface layer. The surface electric field is proportional to the slope of the bands at the surface, and is reduced in this structure.

8.2.10 Inversion Layer Electron Mobility

8.2.10.1 Mechanisms Affecting Inversion Layer Mobility

In the development of the current–voltage characteristics of the MOSFET in Section 8.2.3, we denoted the mobility of electrons in the inversion layer by the symbol c08-math-0502. At inversion biases, electrons are confined to the oxide/semiconductor interface by the strong band bending in the semiconductor. The mobility of electrons in the inversion layer is lower than in the bulk semiconductor due to increased scattering at the oxide/semiconductor interface. Moreover, the mobility decreases with increasing gate voltage, since higher gate voltages increase the electric field confining electrons to the interface, resulting in increased scattering. The gate voltage dependence can be described by the empirical equation

where c08-math-0504 is the peak mobility at threshold and c08-math-0505 is a parameter characterizing the rate of decrease with field.

In silicon MOSFETs, the inversion layer mobility near threshold is about half the bulk mobility, but in 4H-SiC the inversion layer mobility on the silicon face is only about 5–10% of the bulk mobility. This impacts the performance of power MOSFETs by increasing the specific on-resistance, as shown by Equation 8.52. The low inversion layer mobility is a major limitation for MOSFETs and IGBTs in 4H-SiC, and inversion layer transport is the subject of intensive research.

Electron mobility in the inversion layer is limited by several scattering mechanisms, including surface phonon scattering, Coulomb scattering by fixed charges and charged interface states, and surface roughness scattering due to the structural and stoichiometric disorder at the interface. In addition, inversion electrons are also subject to the same scattering mechanisms as electrons in the bulk semiconductor. Mobility is inversely proportional to the total scattering rate, and assuming that scattering rates from the different processes add, the resultant mobility can be expressed using Matthiesson's rule as

where c08-math-0507 is the mobility of electrons in the bulk semiconductor, c08-math-0508 is the mobility due to surface phonon scattering, c08-math-0509 is the mobility due to Coulomb scattering, and c08-math-0510 is the mobility due to surface roughness scattering. The inversion layer mobility is therefore determined by the interplay of several scattering mechanisms whose magnitudes are affected by device processing and operating conditions. Many of these dependences can be related to the effective normal field, defined as the electric field in the semiconductor normal to the surface, evaluated at the centroid of the inversion layer [5]. Before considering the scattering mechanisms individually, we will examine how the effective normal field depends on parameters such as gate voltage, doping, and temperature.

In silicon, inversion layer mobility follows a universal curve when plotted against the effective normal field c08-math-0511, which can be written [5]

Here c08-math-0513 is the sheet charge density in the inversion layer and c08-math-0514 is the charge per unit area in the semiconductor depletion region. In an operating transistor, both c08-math-0515 and c08-math-0516 are functions of position c08-math-0517 along the channel. Using Equations 8.29, and 8.34, the depletion charge can be written

where c08-math-0519 is given by Equation 8.21 and c08-math-0520 by Equation 8.28. The quasi-Fermi level splitting c08-math-0521 is a function of position along the channel, going from c08-math-0522 at the source to c08-math-0523 at the drain. Under delta-depletion electrostatics, the charge in the semiconductor depletion region does not depend on gate voltage in inversion, as shown by Equation 8.64. A more exact analysis would reveal a weak dependence that we will neglect in the present discussion.

The inversion charge c08-math-0524 can be obtained from the gate voltage/surface potential relation of Equation 8.27 after appropriate modifications. Equation 8.27 is derived assuming depletion-region biasing with no inversion layer present. To apply Equation 8.27 in inversion, we must replace c08-math-0525 with an “effective” gate voltage that includes the screening effect of the inversion charge. Performing this modification and setting the surface potential in inversion c08-math-0526, we obtain

8.65 equation

Solving for c08-math-0528, and expanding the c08-math-0529 term to include charge in interface states c08-math-0530, we can write

We can apply Equations 8.64 and 8.66 at the source by setting c08-math-0532. Then using Equation 8.59, the inversion charge at the source reduces to the simple expression c08-math-0533, showing that the inversion charge increases linearly with gate voltage once c08-math-0534 exceeds c08-math-0535.

c08-math-0536 and c08-math-0537 depend on temperature through two effects. As shown in Figure A.3, the Fermi potential c08-math-0538 decreases with temperature, that is, the Fermi level moves closer to midgap as the temperature is raised. This means that less band bending is required to reach inversion, and this is represented in Equations 8.59 and 8.66 by the c08-math-0539 terms. The c08-math-0540 term in these equations is the charge in interface states, which depends on the position of the Fermi level at the surface. For an n-channel MOSFET in inversion, the Fermi level lies above midgap at the surface by an amount equal to c08-math-0541. As the temperature is raised and the Fermi level moves closer to midgap, the negative charge in interface states decreases and c08-math-0542 becomes less negative. For an n-channel MOSFET, both these effects make c08-math-0543 more negative and c08-math-0544 less positive as the temperature is raised. In contrast, c08-math-0545 becomes less negative with temperature, since c08-math-0546 decreases with temperature (Figure A.3). Since c08-math-0547 and c08-math-0548 change in opposite directions as the temperature is raised, it often turns out that the effective normal field at a fixed gate voltage is only a weak function of temperature, as can be verified using the above equations.

Equations 8.63 and 8.64 indicate that samples with heavier substrate doping (larger c08-math-0549 and c08-math-0550) have a higher normal field at the same inversion charge density, and this is expected to increase scattering and reduce mobility by confining electrons closer to the interface. In addition, in heavier-doped samples the Fermi level lies closer to the conduction band in inversion, so there is more negative charge in interface states, and this can increase Coulomb scattering.

We will now consider each of the four scattering mechanisms in Equation 8.62. The mobility due to scattering in the bulk semiconductor depends on doping and temperature, and can be described by an equation of the form [6]

8.67 equation

where c08-math-0552 is the peak mobility at 300 K, c08-math-0553 is absolute temperature, c08-math-0554 and c08-math-0555 are the ionized dopant concentrations (which depend upon temperature), and c08-math-0556 and c08-math-0557 are constants. Values for these parameters in 4H-SiC are given in Table 8.1, and Appendix A gives equations for c08-math-0558 and c08-math-0559 as a function of temperature and doping. For normal operating conditions, the bulk mobility term is not the limiting factor in SiC inversion layer mobility.

Table 8.1 Parameters of the inversion layer mobility model, Equations 8.70.

Parameter Value Units References
c08-math-0560 1141 c08-math-0561 [8, 9]
c08-math-0562 c08-math-0563 c08-math-0564
c08-math-0565 2.8
c08-math-0566 0.61
c08-math-0567 c08-math-0568 c08-math-0569 [7]
c08-math-0570 c08-math-0571 c08-math-0572
c08-math-0573 c08-math-0574 c08-math-0575 [6]
c08-math-0576 c08-math-0577 c08-math-0578
c08-math-0579 0.8
c08-math-0580 c08-math-0581 c08-math-0582 [10]

Surface phonon scattering is the deflection of electrons by acoustic phonons at the surface. The phonon-limited mobility can be written [7]

8.68 equation

where c08-math-0584 and c08-math-0585 are parameters evaluated by fitting theory to experiment and have typical values given in Table 8.1. The phonon-limited mobility decreases with increasing normal field (increasing gate voltage) and decreases with increasing temperature, since higher temperatures are associated with larger lattice vibrations. Although acoustic phonon scattering is important in silicon samples, this term is typically not the limiting factor in SiC MOSFETs fabricated to date.

Coulomb scattering is the deflection of electrons by charged centers at the interface, such as fixed charges c08-math-0586 and charged interface states c08-math-0587. The scattering effect is independent of the charge polarity, and depends on the total density of charged centers rather than on the net charge density. A number of different formulations have been proposed for Coulomb scattering, but a commonly accepted form that captures the essential physics is [6]

8.69 equation

where c08-math-0589 is the total density of fixed charges (sum of positive and negative charges), c08-math-0590 is the total density of charged interface states (both positive and negative), c08-math-0591 is the electron density per unit volume at the surface, and c08-math-0592, and c08-math-0593 are parameters with typical values given in Table 8.1. Coulomb scattering limits inversion layer mobility at gate voltages near threshold where c08-math-0594 is small, but drops rapidly at higher gate voltages (higher c08-math-0595 values) due to carrier screening by inversion electrons. In samples with high fixed charge c08-math-0596 or high interface state density c08-math-0597, Coulomb scattering can limit c08-math-0598 to single digits, but improvements in oxidation and anneal processes have reduced both c08-math-0599 and c08-math-0600 to the point that Coulomb scattering is comparable to surface roughness scattering in limiting inversion layer mobility.

Surface roughness scattering is the deflection of electrons by structural defects at the interface. In silicon, surface roughness scattering is attributed to surface steps and residual unoxidized silicon particles at the interface. However, in SiC the situation is more complex, and there is evidence for a transition layer at the interface where the chemical composition changes gradually from pure SiC to pure c08-math-0601 over a distance of several nanometers. The transition layer can be seen in high-resolution transmission electron microscope (TEM) images, where it has an apparent width of several nm [11]. Spatially-resolved electron energy loss spectroscopy (EELS) shows a gradual transition of the C/Si ratio, beginning about 3 nm inside the SiC and extending about 5 nm into the c08-math-0602. The surface layer of the SiC also exhibits some structural disorder resulting from the oxidation process, and the first monolayers of c08-math-0603 contain excess carbon [11]. The width of the transition layer is influenced by oxidation and anneal procedures, and samples with thinner transition layers have higher peak inversion mobilities [12].

Surface roughness scattering can be characterized by an expression borrowed from the silicon literature, having the general form [13]

where c08-math-0605 is a parameter that depends on the correlation length and mean height of the assumed roughness. This equation has been applied to 4H-SiC by several authors [7, 14], although the more complex nature of the SiC interface suggests the need for further studies. Nevertheless, the general form of Equation 8.70 indicates the strong dependence of surface roughness scattering on normal field. Roughness scattering limits the mobility of SiC MOSFETs at high normal fields and accounts for the decrease in mobility with gate voltage above threshold. This scattering is essentially independent of temperature, provided the normal field is held constant.

Figure 8.24 illustrates how the four scattering mechanisms vary with inversion charge density (or gate voltage) and temperature. Results are shown for two temperatures, 23 and 344 °C. The arrows indicate the trend of the particular scattering mechanism as temperature is increased. At room temperature, Coulomb scattering limits the mobility in most samples at low electron densities (gate voltages close to threshold) and surface roughness scattering limits at high electron densities (strong inversion). Bulk and surface phonon scattering do not play a significant role at room temperature. As temperature increases, Coulomb and surface roughness scattering decrease, while surface and bulk phonon scattering increase. The overall mobility decreases slightly with temperature, mainly due to the strong increase in bulk phonon scattering. This figure illustrates general trends in a way that is consistent with a broad range of reports in the literature, but it does not describe any particular device. As technology advances, interface state density and surface roughness scattering should continue to decrease, leading to higher mobilities and better MOSFET performance.

c08f024

Figure 8.24 General dependence of surface scattering mechanisms on inversion charge density and temperature. The parameters used to generate these plots were modified slightly from those of Table 8.1 to produce results that are generally consistent with a broad range of reports in the literature.

8.2.10.2 Device-Related Definitions of Inversion Layer Mobility

Having discussed the physical mechanisms governing electron transport in inversion layers, we now wish to consider three different device-related definitions of inversion mobility: effective (or conductivity) mobility c08-math-0606, field-effect mobility c08-math-0607, and Hall mobility c08-math-0608. While closely related, each definition provides a slightly different perspective on conditions and processes within the device.

Equation 8.47 gives the MOSFET current using the “square law” approximation with the source grounded (the effect of a non-zero source voltage is included in Equation 8.44). The effective (or conductivity) mobility is the value deduced by measuring the drain conductance c08-math-0609 at small drain voltages where the c08-math-0610 term in Equation 8.47 can be neglected. Neglecting the c08-math-0611 term, we can write

8.71 equation

Thus,

In practice, c08-math-0614 is calculated as a function of gate voltage from the slope of the c08-math-0615 relation at the origin, c08-math-0616. This mobility should be used for c08-math-0617 in Equations 8.44 and/or 8.47 to calculate the current. Since the mobility varies with gate voltage, c08-math-0618 is usually approximated using the empirical expression in Equation 8.61.

The mobility most often quoted in the literature is the field-effect mobility, obtained from the transconductance measured at a small drain voltage. From Equation 8.47 evaluated at low c08-math-0619 with c08-math-0620, the transconductance may be written

Thus,

From Equation 8.73, the field-effect mobility is related to the effective mobility by

8.75 equation

Since c08-math-0624 decreases with c08-math-0625, the partial derivative is negative and c08-math-0626 at any gate voltage. Figure 8.25 illustrates how c08-math-0627 and c08-math-0628 vary with gate voltage. Near threshold the effective and field-effect mobilities are equal, but c08-math-0629 decreases more rapidly than c08-math-0630 as gate voltage is increased.

c08f025

Figure 8.25 Relationship of effective mobility and field-effect mobility, with values typical of 4H-SiC MOSFETs on the (0001) face after thermal oxidation and post-oxidation anneal in nitric oxide.

It is important to note that using c08-math-0631 instead of c08-math-0632 in Equation 8.44 or 8.47 would underestimate the current. Since c08-math-0633 is a differential mobility, calculating the current based on c08-math-0634 would require integration of the c08-math-0635 versus c08-math-0636 relation. From Equation 8.73, we can write

8.76 equation

Thus the current at a given gate voltage is proportional to the integral of c08-math-0638 up to that gate voltage, and not to the value of c08-math-0639 at that gate voltage. For this reason, any process that increases the peak field-effect mobility in the region just above threshold will increase the current at all gate voltages, even if c08-math-0640 at high gate voltages is not improved.

Finally, we turn to the Hall mobility c08-math-0641. Hall mobility is of interest because it allows us to account for the effect of charge sequestration by interface states. In deriving the current Equations 8.44 and 8.47 we assumed all the additional positive charge placed on the gate beyond threshold is balanced by an equal negative charge induced into the inversion layer, as evident from Equation 8.41. However, if interface states are present, some of the charge placed on the gate beyond threshold is balanced by additional charge in interface states, and this reduces the charge density in the inversion layer. Without knowing the distribution of interface states across the bandgap, it is not possible to calculate how much charge is trapped in interface states and how much remains in the inversion layer. If the actual inversion charge density is less than given by Equation 8.41, measurements of c08-math-0642 and c08-math-0643 using Equations 8.72 and 8.74 will underestimate the true carrier mobility. The effective mobility is the correct mobility to use in the current Equations 8.44 and 8.47, since c08-math-0644 includes the combined effects of inversion layer mobility and charge sequestration by interface states, but the Hall mobility is the correct mobility to use if we are concerned with the true carrier mobility in the inversion layer.

The Hall technique is discussed extensively in the literature, and only an outline will be presented here. When used to study the inversion layer in an MOS transistor, the Hall technique imposes a magnetic field c08-math-0645 perpendicular to the surface with a constant drain current c08-math-0646 flowing along the channel. The magnetic field exerts a Lorentz force on the moving electrons given by

8.77 equation

where c08-math-0648 is the force directed across the width of the channel (c08-math-0649 direction) and c08-math-0650 is the velocity of electrons along the channel in the c08-math-0651 direction. The Lorentz force deflects electrons to one side of the channel, creating a lateral electric field c08-math-0652 perpendicular to the current flow. In steady state the c08-math-0653-directed force due to the lateral electric field exactly balances the oppositely-directed c08-math-0654-directed Lorentz force, so c08-math-0655. The electric field c08-math-0656 produces a lateral potential drop called the Hall voltage, given by

Here, c08-math-0658 is the width of the channel, and we have used the fact that c08-math-0659. The Hall voltage can be measured by high-impedance probes that contact the inversion layer on opposite sides of the channel. Since we know the current density and magnetic field, we can determine the mobile electron density c08-math-0660 using Equation 8.78. The mobility can then be deduced from the measured current, given knowledge of the true carrier density in the channel. The mobility thus determined is known at the Hall mobility c08-math-0661.

A more careful analysis that includes momentum relaxation effects would lead to a slightly modified form for Equation 8.78, namely

8.79 equation

Here c08-math-0663 is the Hall factor given by c08-math-0664, where c08-math-0665 is the mean time between electron scattering events. In 4H-SiC the Hall factor at room temperature typically lies in the range 0.96–0.99 [15], and in experimental work it is common to simply set c08-math-0666.

Since Hall measurements are based on the actual mobile carrier density, the Hall mobility represents the true mobility of electrons in the inversion layer, and c08-math-0667 is invariably higher than c08-math-0668.

8.2.10.3 Experimental Results on Inversion Layer Mobility in 4H-SiC

Figure 8.26 shows field-effect mobility measured as a function of temperature for a 4H-SiC MOSFET on a c08-math-0669 p-type epilayer doped c08-math-0670 with aluminum [16]. The c08-math-0671 (0001) substrate is grown 8° off-axis. A 54 nm gate oxide was formed by pyrogenic oxidation at 1150 °C, followed by 30 min in situ argon anneal, a 950 °C re-oxidation anneal for 2 h, and a nitric oxide (NO) post-oxidation anneal at 1175 °C for 2 h. The final NO anneal is used to reduce the interface state density in the upper half of the bandgap [17]. Polysilicon gates were deposited at 625 °C and doped at 900 °C. At room temperature, the peak mobility is about c08-math-0672, decreasing to about c08-math-0673 at 15 V. As temperature is increased, the peak low-field mobility increases to approximately c08-math-0674 at 167 °C, consistent with Coulomb scattering, then decreases to about c08-math-0675 at 344 °C. At higher fields the mobility decreases monotonically with temperature. The decrease in mobility with temperature is suggestive of bulk phonon scattering, as can be inferred from the curves in Figure 8.24. At 344 °C the field-effect mobility has dropped to about 80% of its room temperature value over much of the gate voltage range.

c08f026

Figure 8.26 Field-effect mobility as a function of temperature for a 4H-SiC MOSFET on a p-type epilayer ([16] reproduced with permission from IEEE).

Figure 8.27 shows Hall mobility as a function of sheet carrier density for an n-channel MOSFET on a p-type epilayer doped c08-math-0676 with aluminum [18]. The substrate is (0001) 4H-SiC, cut 4° off-axis. The oxidation was performed in dry oxygen at 1175 °C, followed by a wet re-oxidation anneal at 950 °C and an 1175 °C post-oxidation anneal in NO for 2 h, resulting in a 53 nm oxide. This is a similar oxidation process to the MOSFET of Figure 8.26, but with a slightly lower epilayer doping. Hall mobility increases with temperature below 20 °C, suggesting that transport at these temperatures is limited by Coulomb scattering. Above room temperature the mobility increases only slightly with temperature, suggesting that transport here is dominated by surface roughness scattering. These conclusions are consistent with the trends illustrated in Figure 8.24.

c08f027

Figure 8.27 Hall mobility of inversion electrons as a function of effective normal field at several temperatures. The sample was prepared by a similar oxidation process to the MOSFET in Figure 8.26 ([18] reproduced with permission from Trans Tech Publications).

The MOSFET process described above has been investigated by several groups, all of which report increased mobilities as a result of the NO anneal. Recently, however, a number of alternative oxidation processes have been explored, some of which have produced even greater improvements in mobility. We will briefly discuss two of these processes next.

Post-oxidation annealing in phosphorus has been shown to reduce the interface state density and increase the mobility, even compared to the standard NO process. Figure 8.28 shows measured interface state density and field-effect mobility for several 4H-SiC MOSFETs [19]. The MOSFETs were formed on p-type epilayers doped c08-math-0677 with aluminum, and interface state density was measured on MOS capacitors on n-type epilayers doped c08-math-0678 with nitrogen. The substrate is c08-math-0679 (0001), grown 4° off-axis. A 56 nm gate oxide was formed by dry oxidation at 1000 °C, followed by a post-oxidation anneal in c08-math-0680 at 900, 950, or 1000 °C for 10 min and a 30 min nitrogen anneal at the same temperature. Aluminum was evaporated to form the gate electrodes. Other samples were fabricated by dry oxidation at 1000 °C followed by a 1250 °C nitric oxide (NO) anneal for 90 min. As seen in the figure, the 900 °C phosphorus anneal did not reduce the interface state density, but anneals at 950 and 1000 °C decreased the interface state density below that of the NO sample in the upper half of the bandgap. The field-effect mobility of the 1000 °C c08-math-0681 sample is much higher than the 1250 °C NO sample, reaching a peak value of c08-math-0682 before dropping to c08-math-0683 at 20 V. If this result is compared to the MOSFET of Figure 8.26, the mobility improvement is not as great, but is still about a factor of 2.

c08f028

Figure 8.28 (a) Interface state density and (b) field-effect mobility for dry oxidized samples with no post-oxidation anneal, with a post-oxidation anneal in nitric oxide, and with a post-oxidation anneal in c08-math-0684. The large increase in mobility for the c08-math-0685 sample is correlated with a reduction in interface state density ([19] reproduced with permission from IEEE).

The improved mobility is explained as a reduction in Coulomb scattering by charged interface states, but the phosphorus process may also introduce a shallow n-type layer at the surface of the SiC. This seems plausible, since phosphorus can be activated as a dopant in 4H-SiC at normal oxidation temperatures. This would produce a situation similar to the doped-channel FET discussed in Section 8.2.9. As shown in Figure 8.23, surface doping reduces the surface normal field at a given inversion electron density, which would lead to higher mobility. However, phosphorus annealing also converts the oxide to phosphosilicate glass (PSG), a polar material having piezoelectric properties that introduce an instability in the threshold voltage [20]. Phosphorus atoms are distributed throughout the gate insulator, and they act as electron traps that introduce a positive threshold shift at high electric fields [21]. Instabilities of this type would make it difficult to use this process in a production environment.

Several groups have explored inversion layer transport on alternative crystal faces such as the c08-math-0686 a-face, where mobilities are typically higher than on the (0001) silicon face. These results are important for devices such as the UMOSFET (or UMOS IGBT) whose inversion layers are on the a-face. Figure 8.29 shows measured interface state density and field-effect mobility for 4H-SiC MOSFETs on both the silicon face and the c08-math-0687 a-face [22]. Planar MOSFETs were formed on p-type epilayers doped c08-math-0688 with aluminum, and interface state density was measured on MOS capacitors on n-type epilayers doped c08-math-0689 with nitrogen. All samples were oxidized in dry oxygen at 1150 °C. Some samples received a post-oxidation anneal in NO at 1175 °C for 2 h, and others received a phosphorus post-oxidation anneal at 1000 °C for 4 h using a planar diffusion source. On a given crystal face, the field-effect mobility is c08-math-0690 higher with the phosphorus anneal compared to the NO anneal. For the same post-oxidation process, the mobility is also c08-math-0691 higher on the c08-math-0692 face than on the silicon face. The highest mobility is obtained with the phosphorus anneal on the c08-math-0693 face.

c08f029

Figure 8.29 (a) Interface state density and (b) field-effect mobility for dry oxidized samples on two crystal orientations, the (0001) silicon face and the c08-math-0694 a-face. Mobility is shown for both nitrogen and phosphorus post-oxidation anneals, and interface state density is shown for unannealed, nitrogen-annealed, and phosphorus-annealed samples ([22] reproduced with permission from IEEE).

With the phosphorus anneal on the silicon face, the higher mobility is correlated with a lower interface state density, suggesting the mobility is limited by Coulomb scattering, at least at biases near threshold. However, on the c08-math-0695 face the opposite is true: higher mobilities are obtained with the phosphorus anneal, but the interface state density is similar on both faces. Comparing phosphorus anneals on the two faces, a higher mobility is obtained on the c08-math-0696 face even though that face has a higher interface state density. Two possibilities present themselves. Yoshioka et al. [23] have reported that certain annealing procedures create interface states with large capture cross sections whose frequency response is so fast that they are not detected with conventional CV methods, so the conclusion that the silicon face has lower total interface state density may be incorrect. On the other hand, if phosphorus achieves high mobility by creating a thin n-doped layer at the surface, this doping process could be more effective on the c08-math-0697 face than on the silicon face. These issues can only be resolved by further experiments and the reader is advised to consult the literature for the latest information.

An intriguing result is the high mobility observed in samples oxidized in an alumina furnace tube [24], as shown in Figure 8.30 [25]. The high mobility is correlated with the presence and position of sodium ions in the oxide, introduced from impurities in the alumina tube. These ions are mobile in c08-math-0698 at elevated temperatures, and can be drifted toward the interface or toward the gate by application of a positive or negative gate voltage. The highest mobilities are observed when the sodium ions are close to the interface, the “initial” and “recovered” curves in Figure 8.30. When the sodium ions are drifted to the gate by negative bias-temperature stress, the threshold voltage increases, and the mobility decreases. Measurements of interface state density on n-type MOS capacitors indicate that the position of the sodium ions has no effect on interface state density [25], suggesting that a reduction of Coulomb scattering is not the mechanism responsible for the mobility improvement. It is also significant that the mobility at high gate fields is not affected by the ions. In other words, sodium enhances the mobility at low gate fields but not at high fields. The physical mechanism responsible for these observations is not understood. As might be expected, the instability in threshold voltage associated with sodium precludes the use of this process in practical devices.

c08f030

Figure 8.30 Field-effect mobility of a MOSFET whose oxide is contaminated with sodium ions. Negative bias-temperature stress (BTS) draws the positive sodium ions to the gate, where they have no effect on the electrostatics. This shifts the threshold positive and reduces the mobility. After the stress is removed, the threshold and mobility gradually recover to their pre-stress values ([25] reproduced with permission from Trans Tech Publications).

Even though both phosphorus and sodium processes introduce threshold instabilities, the results are still encouraging because they show that significantly higher mobilities are possible in 4H-SiC MOS devices. The goal now is to understand the mechanisms for the mobility improvement and develop processes that deliver the improvement without introducing instabilities.

8.2.11 Oxide Reliability

Two of the main advantages of SiC compared to silicon are its higher bandgap energy and higher critical field for avalanche breakdown. The critical field in 4H-SiC is 6–7 times higher than in silicon (Figure 10.5) and, as discussed in Chapter 7, the on-resistance of SiC unipolar devices is about 400 times lower than silicon devices of the same blocking voltage. However, the higher critical field means that gate oxides in SiC MOS devices may be subjected to higher fields than oxides in silicon devices. The oxide field is related to the surface field in the semiconductor by Gauss' law, so we can write

8.80 equation

where c08-math-0700 is the perpendicular component of the semiconductor field at the surface. In SiC, c08-math-0701 (and therefore c08-math-0702) can be 6–7 times higher than in silicon. The oxide field may be increased even further by fixed charges c08-math-0703 and interface trapped charges c08-math-0704. In addition, the barrier heights for electron and hole injection into c08-math-0705 are lower for SiC, as shown in Figure 8.21. The higher oxide field and lower barrier height raise concerns for the long term reliability of SiC MOS devices, particularly at high temperatures.

The literature contains very few careful studies of SiC MOS reliability. One reason is the large number of samples (20–50) and long measurement times (weeks to months) needed to obtain an adequate data set. Most studies make use of constant-voltage stress at elevated temperatures. The procedure is to subject a number of identical devices, either MOS capacitors or MOSFETs, to accelerated stress conditions (high oxide fields and high temperatures) and monitor the leakage currents through the gate oxides. A certain current density is chosen as indicative of oxide failure, and a set of devices is stressed simultaneously while the gate currents are monitored. As devices fail, their failure times are recorded and the test continues until the entire set of devices has failed. The failure times are plotted on a Weibull plot, where it is possible to distinguish between extrinsic and intrinsic failures. In this context, “extrinsic” refers to early failures due to oxide defects, and “intrinsic” refers to failure of an intrinsically good oxide. Since the extrinsic failures do not represent the fundamental properties of the oxide, they are eliminated from the distribution and the mean-time-before-failure c08-math-0706 or the 63% failure time c08-math-0707 of the remaining intrinsic set is determined. Constant-voltage stress is considered fairly representative of actual operating conditions. A quicker procedure, known as constant-current stress, forces a constant current through the oxide and measures the total charge-to-breakdown, or c08-math-0708. The discussion below deals only with constant-voltage stress.

Figure 8.31 shows 63% failure times as a function of oxide field for 4H-SiC MOS capacitors on n-type epilayers from two vendors [26]. As is the case in silicon [27], the time-to-failure increases exponentially as the oxide field is reduced, and a constant field acceleration factor c08-math-0709 is observed. At 225 °C, extrapolation to lower fields predicts a c08-math-0710 of 100 years if the oxide field is kept below c08-math-0711. The oxide field must be kept below c08-math-0712 if a c08-math-0713 of 100 years is to be obtained at 375 °C. Although 100 year lifetimes may seem excessive, this is the time for 63% of the samples to fail. Failure times for lower failure percentages, say 10%, can be determined by examining the failure distributions in the Weibull plots at each field.

c08f031

Figure 8.31 63% failure time as a function of oxide field for 4H-SiC MOS capacitors on n-type epilayers at two temperatures ([26] reproduced with permission from IEEE).

Figure 8.32 shows c08-math-0714 failure time versus oxide field for 4H-SiC power DMOSFETs at three temperatures [28]. The n-channel DMOSFETs are formed on implanted p-type base regions, and a portion of the gate extends over implanted c08-math-0715 source regions. Oxides over heavily-implanted regions are not as robust as those over epitaxial layers, and the c08-math-0716 times for the DMOSFET are shorter than the MOS capacitors of Figure 8.31. Extrapolation of the high-field data indicates that a c08-math-0717 of 100 years can be realized at 175 °C by keeping the oxide field below c08-math-0718. The same lifetime can be obtained at 275 °C by keeping the oxide field below c08-math-0719, and at 300 °C by keeping the oxide field below c08-math-0720.

c08f032

Figure 8.32 63% failure time as a function of oxide field for 4H-SiC power DMOSFETs at three temperatures ([28] reproduced with permission from Trans Tech Publications).

The data in Figure 8.32 also illustrate the danger in extrapolating high-field data to lower fields. At 275 and 300 °C the field acceleration factor c08-math-0721 is not independent of field, but exhibits a break at an oxide field around c08-math-0722. If extrapolations were made using only data above c08-math-0723, unrealistically long lifetimes would be predicted at normal operating fields. This brings into question the practice of predicting low-field reliability by extrapolation from high-field data. However, it seems safe to assume that the low-field reliability will not be better than predicted by extrapolating from high-field data. The severity of the change in field acceleration factor in Figure 8.32 decreases as temperature is reduced, and at 175 °C it appears that extrapolation using the high-field acceleration factor is reasonable.

Finally, we should point out that the operational lifetime of the MOSFET may be limited not by catastrophic oxide failure, but by a gradual drift in parameters such as threshold voltage or on-resistance. Unfortunately, no studies of long-term parameter drift have been published for SiC devices, leaving the question of parameter stability largely unanswered.

8.2.12 MOSFET Transient Response

One of the motivations for using MOSFETs in power switching systems is their high switching frequency and low switching loss. This allows the entire system to operate at a high frequency, leading to significant reductions in the volume and weight of passive components, such as transformers and filter capacitors, that often dominate system cost. In this section we will consider the transient response of the power MOSFET in some detail.

The switching transients of the power MOSFET can be analyzed using the clamped inductive load circuit of Figure 8.33. The inductor is representative of the windings of an electric motor, and its inductance is assumed large enough that the current c08-math-0724 cannot change appreciably during the switching transient. c08-math-0725 represents the internal resistance of the bond wires and the distributed gate resistance of the MOSFET. c08-math-0726 and c08-math-0727 are the gate-source and gate-drain capacitances of the MOSFET, c08-math-0728 is the source resistance, and c08-math-0729 is the lumped drain resistance of the MOSFET.

c08f033

Figure 8.33 Equivalent circuit used for the transient analysis of a MOSFET driving a clamped inductive load. The dashed box encloses the MOSFET and its internal parasitic elements.

Since the p-type body regions of power MOSFETs are not lightly doped, the c08-math-0734 terms in Equations 8.44 and 8.48 cannot be ignored, and the simplified Equations 8.47 and 8.49 will overestimate the current and underestimate the switching time. However, retaining all the c08-math-0735 terms would make the mathematics very cumbersome. Therefore, to illustrate the main features of the switching transient, we will utilize the approximate Equations 8.47 and 8.49 and assume the source resistance c08-math-0736 is negligible. This will not provide numerically precise answers, but will capture the qualitative features of the transient. A more quantitative analysis can be performed using the full MOSFET equations and a nonlinear circuit simulator such as c08-math-0737.

The turn-on transient will be considered first. Prior to c08-math-0738 the MOSFET is off, and the inductor current circulates through the clamping diode. If we neglect the forward drop of the diode, which is likely to be a SiC junction-barrier Schottky (JBS) or merged pin-Schottky (MPS) diode, the drain voltage of the MOSFET is the supply voltage c08-math-0739. The turn-on process can be divided into four phases, as illustrated in Figure 8.34.

c08f034

Figure 8.34 Drain characteristics of the MOSFET to be analyzed, with c08-math-0730, and c08-math-0731. In these calculations, c08-math-0732, and c08-math-0733 switches between 0 and 20 V.

8.2.12.1 Turn-On, 0 < t < t1

At c08-math-0740 the voltage at the gate terminal is stepped to c08-math-0741, but the MOSFET does not turn on until c08-math-0742 has charged to c08-math-0743 through the RC input circuit. The gate voltage during the charging process can be written

where c08-math-0745 is given by

Setting c08-math-0747 allows us to calculate c08-math-0748,

8.83 equation

The gate current during this phase is of interest, and can be written

8.84 equation

8.2.12.2 Turn-On, t1 < t < t2

At time c08-math-0751 the MOSFET turns on and its drain current increases, but the clamping diode remains forward biased until the drain current through the MOSFET equals the inductive load current c08-math-0752. As long as the diode is forward biased, the drain-source voltage of the MOSFET remains at c08-math-0753 and the gate-source voltage continues to increase according to Equation 8.81. Since c08-math-0754, the MOSFET is in its saturation region and the drain current can be described by Equation 8.47 with c08-math-0755 set equal to c08-math-0756, namely

The drain current will increase according to Equations 8.85 and 8.81 until time c08-math-0758 when c08-math-0759, at which point the MOSFET is carrying all the load current and the diode becomes reverse biased. The trajectory between times c08-math-0760 and c08-math-0761 is illustrated on the c08-math-0762 characteristics of the MOSFET in Figure 8.34. Time c08-math-0763 can be calculated by setting c08-math-0764 in Equation 8.85 to obtain c08-math-0765, and inserting this into Equation 8.81 to solve for c08-math-0766. The result is

8.86 equation

8.2.12.3 Turn-On, t2 < t < t3

When the diode becomes reverse biased, the drain voltage of the MOSFET is no longer clamped at c08-math-0768, and is free to decrease. However, the drain current is now held fixed at c08-math-0769 by the inductor. With the drain current constant, the gate-source voltage cannot change, as shown by Equation 8.85. Since c08-math-0770 has not yet reached c08-math-0771, gate current continues to flow through c08-math-0772 and passes as displacement current through c08-math-0773. (Although c08-math-0774 is constant, c08-math-0775 is now falling, and this draws displacement current through c08-math-0776. In claiming that c08-math-0777 is constant, we are assuming this displacement current is small compared to c08-math-0778.) The gate current can now be written

8.87 equation

Writing c08-math-0780 and setting c08-math-0781 yields

Since c08-math-0783 and c08-math-0784 are constant during this period, Equation 8.88 tells us that the drain voltage decreases linearly with time. The drain voltage transient can be calculated by integrating Equation 8.88,

The above analysis holds so long as the MOSFET remains in its saturation region, where the drain current is given by Equation 8.85. When c08-math-0786 falls below c08-math-0787, the MOSFET enters its quasi-linear region and c08-math-0788 must be calculated using Equation 8.47. This marks the end of the third portion of the transient, and the time c08-math-0789 can be found by setting c08-math-0790 equal to c08-math-0791,

8.90 equation

8.2.12.4 Turn-On, t > t3

For c08-math-0793 the MOSFET is in its quasi-linear region. The gate current can be written

8.91 equation

Since c08-math-0795 is now c08-math-0796, we can set c08-math-0797, so

8.92 equation

Cross-multiplying,

Integrating Equation 8.93 from c08-math-0800 to c08-math-0801 yields

8.94 equation

where we are reminded that c08-math-0803. During the period c08-math-0804 the drain voltage decreases toward the steady-state drain–source voltage of the MOSFET. Setting c08-math-0805 in Equation 8.47 equal to c08-math-0806 and solving for c08-math-0807 yields

As the device approaches steady state, c08-math-0813 approaches c08-math-0814. Inserting c08-math-0815 into Equation 8.95 yields

The total gate charge during turn-on is found by setting c08-math-0817. c08-math-0818 goes from zero to c08-math-0819 during the transient, while c08-math-0820 goes from c08-math-0821 at c08-math-0822 to c08-math-0823 at c08-math-0824. Using Equation 8.96 for c08-math-0825 yields

Figure 8.35a–c show gate voltage, drain current, and drain voltage for the MOSFET of Figure 8.34 as calculated using the above equations with parameters typical of a modern commercial power DMOSFET. In these calculations, the supply voltage c08-math-0827, the terminal gate voltage c08-math-0828, and the inductive load current c08-math-0829 A. The time c08-math-0830 for c08-math-0831 to reach threshold is short, only 0.52 ns. The MOSFET drain current then rises according to Equation 8.85 until reaching the 20 A load current at time c08-math-0832 ns. At this point the diode becomes reverse biased and the MOSFET drain voltage decreases linearly according to Equation 8.89 until time c08-math-0833 ns. During this period c08-math-0834 remains constant at 11.8 V, c08-math-0835 is constant at 1.78 A, and c08-math-0836 is constant at 20 A. At time c08-math-0837 the drain voltage has reached c08-math-0838 and the MOSFET enters its quasi-linear region. From this point, c08-math-0839 declines according to Equation 8.95 to a final value of c08-math-0840.

c08f035

Figure 8.35 Transient waveforms during the turn-on transient, assuming c08-math-0809. (a) Internal gate-source voltage c08-math-0810, (b) drain current c08-math-0811, (c) drain-source voltage c08-math-0812, and (d) instantaneous power dissipation within the MOSFET.

The instantaneous power dissipated during the turn-on transient is given by

8.98 equation

The instantaneous power dissipation is plotted in Figure 8.35d, and the integral of this waveform gives the turn-on energy, c08-math-0842 in this example. From Equation 8.97, the gate charge during the turn-on transient is 36.6 nC.

The turn-off transient follows the same trajectory as the turn-on transient, but in the opposite direction, as illustrated in Figure 8.36. Prior to the start of the transient at c08-math-0843, the MOSFET is in its conducting state with a very low forward voltage drop. The clamping diode is reverse biased, and all the load current flows through the MOSFET.

c08f036

Figure 8.36 Idealized switching trajectory of the MOSFET during turn-off.

8.2.12.5 Turn-Off, 0 < t < t4

At c08-math-0844 the voltage at the gate terminal is switched to zero, but the MOSFET does not turn off immediately because the RC input circuit must discharge. As c08-math-0845 decreases, c08-math-0846 must increase to keep the drain current at a constant value equal to c08-math-0847, but the change in c08-math-0848 during this period is small compared to the change in c08-math-0849, so we can write the gate voltage as

where c08-math-0851 is the time constant of the gate circuit, given by Equation 8.82. However, as c08-math-0852 increases, it eventually reaches c08-math-0853 and the MOSFET enters saturation. At this point we can write

8.100 equation

Defining this point as c08-math-0855 and solving for c08-math-0856,

Inserting Equation 8.101 into Equation 8.99 gives

8.102 equation

While the MOSFET is in its quasi-linear region, the drain current is given by

8.103 equation

Solving for c08-math-0860, we can write

8.104 equation

with c08-math-0862 given by Equation 8.99.

8.2.12.6 Turn-Off, t4 < t < t5

Between times c08-math-0863 and c08-math-0864 the MOSFET is in saturation and the drain current is fixed at c08-math-0865. Since the drain current remains constant, the gate voltage cannot change, and remains at the value given by Equation 8.101. The gate current can be written

8.105 equation

Writing c08-math-0867 and setting c08-math-0868 yields

Equation 8.106 tells us that the drain voltage increases linearly with time,

At c08-math-0871, so from Equation 8.107 we have

8.108 equation

where we have made use of the fact that c08-math-0873.

8.2.12.7 Turn-Off, t5 < t < t6

As c08-math-0877 tries to rise above c08-math-0878, the clamping diode becomes forward biased and holds c08-math-0879 at c08-math-0880 plus a diode drop (which we will neglect). The diode carries an increasing fraction of the load current, and the MOSFET drain current decreases toward zero. Since c08-math-0881 decreases with c08-math-0882 held constant, c08-math-0883 must also decrease. Writing the equation for the gate current,

Solving Equation 8.109 for c08-math-0885,

At c08-math-0887 reaches c08-math-0888 and the MOSFET turns off. Setting c08-math-0889 in Equation 8.110, we find that

8.111 equation

During the period from c08-math-0891 to c08-math-0892, the MOSFET is in its saturation region and c08-math-0893 is given by Equation 8.85, with c08-math-0894 given by Equation 8.110.

8.2.12.8 Turn-Off, t > t6

For times c08-math-0895 the MOSFET is off, and the gate voltage c08-math-0896 decreases toward zero according to Equation 8.110.

The gate charge c08-math-0897 removed by the input circuit during the turn-off transient is precisely the same as the charge supplied during the turn-on transient, since at the end of the turn-off process both c08-math-0898 and c08-math-0899 have returned to their original charge states.

Figure 8.37a–c show gate voltage, drain current, and drain voltage for the MOSFET of Figure 8.36. The time c08-math-0900 for c08-math-0901 to rise to c08-math-0902 is 2.33 ns. During this time the change in c08-math-0903, 9.63 V, is almost imperceptible on the scale of Figure 8.37c. The MOSFET enters saturation at c08-math-0904 and the drain current remains fixed at the load current, 20 A. According to Equation 8.85, c08-math-0905 must also remain constant. However, c08-math-0906 can rise because c08-math-0907 is independent of c08-math-0908 in saturation. The increase in c08-math-0909 is just sufficient to create the displacement current in c08-math-0910 required to satisfy Equation 8.106. The drain voltage reaches the supply voltage at time c08-math-0911 ns. At this point the clamping diode becomes forward biased and holds c08-math-0912 constant at c08-math-0913. After c08-math-0914 the diode carries an increasing fraction of the load current, and c08-math-0915 decreases. The drain current reaches zero at c08-math-0916 ns, when c08-math-0917 has fallen to c08-math-0918. The gate voltage continues to decrease as the input circuit completes its discharge. Figure 8.37d shows the instantaneous power during the turn-off transient, and the integral of this waveform is the turn-off energy, c08-math-0919.

c08f037

Figure 8.37 Transient waveforms during the turn-off transient. (a) Internal gate-source voltage c08-math-0874, (b) drain current c08-math-0875, (c) drain-source voltage c08-math-0876, and (d) instantaneous power dissipation within the MOSFET.

References

  1. [1] Palmour, J.W., Edmond, J.A., Kong, H.S., and Carter, C. (1993) 6H-SiC power devices for aerospace applications. Proceedings of the 28th Intersociety Energy Conversion Conference, pp. 1.249–1.254.
  2. [2] Shenoy, J.N., Melloch, M.R., and Cooper, J.A. (1996) High-voltage double-implanted power MOS transistors in 6H-SiC. IEEE Device Research Conference, Santa Barbara, CA.
  3. [3] Saha, A. and Cooper, J.A. (2007) A 1200 V 4H-SiC power DMOSFET optimized for low on-resistance. IEEE Trans. Electron. Devices, 54 (10), 2786–2791.
  4. [4] Tan, J., Cooper, J.A. and Melloch, M.R. (1998) High-voltage accumulation-layer UMOSFETs on 4H-SiC. IEEE Electron. Device Lett., 19 (12), 487–489.
  5. [5] Sabnis, A.G. and Clemens, J.T. (1979) Characterization of the electron mobility in the inverted <100> Si surface. IEEE International Electron Devices Meeting Technical Digest, pp. 18–21.
  6. [6] Powell, S.K., Goldsman, N., McGarrity, J.M. et al. (2002) Physics-based numerical modeling and characterization of 6H-silicon carbide metal-oxide-semiconductor field-effect transistors. J. Appl. Phys., 92 (7), 4053–4061.
  7. [7] Potbhare, S., Goldsman, N., Pennington, G. et al. (2006) Numerical and experimental characterization of 4H-silicon carbide lateral metal-oxide-semiconductor field-effect transistor. J. Appl. Phys., 100 (4), 044515.1–044515.8.
  8. [8] Schaffer, W.J., Negley, G.H., Irvine, K.J. and Palmour, J.W. (1994) Conductivity anisotropy in epitaxial 6H and 4H-SiC, in Diamond, SiC, and Nitride Wide-Bandgap Semiconductors, Materials Research Society Proceedings, (eds C.H. Carter Jr., G. Gildenblatt, S. Nakamura and R.J. Nemanich), MRS, vol. 399, pp. 595–600.
  9. [9] Ruff, M., Mitlehner, H. and Helbig, R. (1994) SiC devices: physics and numerical simulaton. IEEE Trans. Electron. Devices, 41 (6), 1040–1054.
  10. [10] Tilak, V. and Matocha, K. (2007) Electron-scattering mechanisms in heavily doped silicon carbide MOSFET inversion layers. IEEE Trans. Electron. Devices, 54 (11), 2823–2829.
  11. [11] Taillon, J. A., Yang, J. H., Ahyi, C. A. et al. (2013) Systematic structural and chemical characterization of the transition layer at the interface of NO-annealed c08-math-0920 metal-oxide-semiconductor field-effect transistors. J. Appl. Phys., 113 (4) 044517.1–044517.6.
  12. [12] Biggerstaff, T.L., Reynolds, C.L. Jr., Zheleva, T. et al. (2009) Relationship between c08-math-0921 transition layer thickness and mobility. Appl. Phys. Lett., 95 (3), 032108.1–032108.3.
  13. [13] Zeng, Y.A., White, M.H. and Das, M.K. (2005) Electron transport modeling in the inversion layers of 4H and 6H-SiC MOSFETs on implanted regions. Solid-State Electron., 49 (6), 1017–1028.
  14. [14] Dhar, S., Haney, S., Cheng, L. et al. (2010) Inversion layer carrier concentration and mobility in 4H-SiC metal-oxide-semiconductor field-effect transistors. J. Appl. Phys., 108 (5), 054509.1–054509.5.
  15. [15] Schmid, F., Krieger, M., Laube, M. et al. (2004) Hall scattering factor of electrons and holes in SiC, in Silicon Carbide—Recent Major Advances (eds W.J. Choyke, H. Matsunami and G. Pensl), Springer, Berlin, pp. 517–536.
  16. [16] Lu, C.Y., Cooper, J.A., Tsuji, T. et al. (2003) Effect of processing variations and ambient temperature on electron mobility at the c08-math-0922 interface. IEEE Trans. Electron. Devices, 50 (7), 1582–1588.
  17. [17] Chung, G.Y., Tin, C.C., Williams, J.R. et al. (2000) Effect of nitric oxide annealing on the interface trap densities near the band edges in the 4H polytype of silicon carbide. Appl. Phys. Lett., 76 (13), 1713–1715.
  18. [18] Dhar, S., Ahyi, A.C., Williams, J.R. et al. (2012) Temperature dependence of inversion layer carrier concentration and Hall mobility in 4H-SiC MOSFETs, in Mater. Sci. Forum, vol. 717–720, pp. 713–716.
  19. [19] Okamoto, D., Yano, H., Harita, K. et al. (2010) Improved inversion channel mobility in 4H-SiC MOSFETs on Si face utilizing phosphorus-doped gate oxide. IEEE Electron. Device Lett., 31 (7), 710–712.
  20. [20] Sharma, Y.K., Ahyi, A.C., Isaacs–Smith, T. et al. (2013) High-mobility stable 4H-SiC MOSFETs using a thin PSG interfacial passivation layer. IEEE Electron. Device Lett., 34 (2), 175–177.
  21. [21] Yano, H., Hatayama, T. and Fuyuki, T. (2012) c08-math-0923 annealing as a new method for improving 4H-SiC MOS device performance. Trans. Electrochem. Soc., 50 (3), 257–265.
  22. [22] Liu, G., Ahyi, A.C., Xu, Y. et al. (2013) Enhanced inversion mobility on c08-math-0924 using phosphorus and nitrogen interface passivation. IEEE Electron. Device Lett., 34 (2), 181–183.
  23. [23] Yoshioka, H., Nakamura, T. and Kimoto, T. (2012) Generation of very fast states by nitridation of the c08-math-0925 interface. J. Appl. Phys., 112 (1), 024520 (6 pages).
  24. [24] Gudjonsson, G., Olafsson, O., Allerstam, F. et al. (2005) High field-effect mobility in n-channel Si-face 4H-SiC MOSFETs with gate oxide grown on aluminum ion-implanted material. IEEE Electron. Device Lett., 26 (2), 96–98.
  25. [25] Sveinbjörnsson, E.O., Gudjonsson, G., Allerstam, F. et al. (2006) High channel mobility 4H-SiC MOSFETs, in Mater. Sci. Forum, vol. 527–529, pp. 961–966.
  26. [26] Yu, L., Cheung, K.P., Campbell, J. et al. (2008) Oxide reliability of SiC MOS devices. 2008 IEEE International Integrated Reliability Workshop Final Report, S. Lake Tahoe, CA, October 12–16, pp. 141–144.
  27. [27] Prendergast, J., Suehle, J., Chaparala, P. et al. (1995) TDDB characterization of thin c08-math-0926 films with bimodal failure populations. IEEE Reliability Physics Symposium, pp. 124–130.
  28. [28] Das, M.K., Haney, S., Richmond, J. et al. (2012) SiC MOSFET reliability update, in Mater. Sci. Forum, vol. 717–720, pp. 1073–1076.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset