5.3 Point Defects in SiC

5.3.1 Major Deep Levels in SiC

5.3.1.1 Intrinsic Defects

The major deep levels that are observed in as-grown n-type and p-type 4H-SiC epitaxial layers are mostly intrinsic defects, and their energy positions in the bandgap are as shown in Figure 4.33. Among them, the c05-math-0496 [152] and EH6/7 c05-math-0497 [153] centers are the dominant thermally stable defects that are most commonly observed with the highest concentration c05-math-0498 in all as-grown epitaxial layers. These levels are also dominant in ion-implanted, plasma-etched, or particle-irradiated SiC, as described in Chapter 6. Although a few hole traps, which may originate from the intrinsic defects, are observed, such as HK4 (or P1) c05-math-0499 [154], they can be annealed out at 1450–1550 °C. Thus, the c05-math-0500 and EH6/7 centers are the most common and important deep levels in SiC. Note that the signals from the EH6 and EH7 centers usually overlap severely in normal DLTS spectra, and these two defect levels are thus often treated as the single EH6/7 center, though the EH7 component is dominant in the “EH6/7 peak”. Laplace-DLTS measurements succeeded in resolving this peak, demonstrating that, in terms of generation upon particle irradiation and thermal stability, there are slight differences between the EH6 and EH7 centers [155]. Here, the features of the c05-math-0501 and EH6/7 centers are summarized.

  1. The c05-math-0509 center exhibits the so-called “negative-U” property, as indicated in Figure 5.42 [156]. One DLTS peak usually reflects the carrier emission from a single deep level unless multiple peaks are overlapping. In the case of the c05-math-0510 center, however, two electrons are emitted simultaneously from the defect level. In the carrier capture process, two electrons are captured almost simultaneously against the Coulomb repulsion between the two electrons, because of the large Jahn–Teller effect [157]. Thus, the intensity of the DLTS signal from the c05-math-0511 center is exactly double that of the real density of the c05-math-0512 center.
  2. The c05-math-0513 center and the EH6/7 center (or at least the EH7 center) are observed with almost identical density in almost all the 4H-SiC samples. Figure 5.43 plots the densities of the c05-math-0514 center versus those of the EH6/7 center for as-grown, electron-irradiated, and annealed 4H-SiC [158, 159]. The densities of both deep levels are varied by changing the epitaxial growth conditions, the electron energy and fluence during irradiation, and the annealing conditions. The c05-math-0515 center density is almost identical to the EH6/7 density for each sample over a wide range from c05-math-0516 to c05-math-0517. This result suggests that the c05-math-0518 center and the EH6/7 (at least the EH7) centers originate from the same defect center but with different charge states.
  3. The origin of the c05-math-0519 center and the EH7 center was unambiguously identified as a carbon vacancy with different charge states, based on the results of theoretical calculations [160] and a comparison study using DLTS and EPR [161, 162]. Figure 5.44a shows the area density of a single-negatively-charged carbon vacancy c05-math-0520 that was determined by EPR versus the area density of the c05-math-0521 center taken from the depth profile that was determined by DLTS measurements [163]. Here, photo-excitation is necessary to observe c05-math-0522, because the carbon vacancy also has a c05-math-0523 nature and stabilizes as c05-math-0524 by capturing two electrons. Because c05-math-0525 is not EPR-active (because it contains no unpaired electrons), the c05-math-0526 density is estimated by producing c05-math-0527 through photo-excitation of n-type SiC. As shown in Figure 5.44a, the c05-math-0528 density is very close to the c05-math-0529 center density for a series of samples. Considering this correspondence and the c05-math-0530 nature described above, the origin of the c05-math-0531 center was identified as the acceptor levels of a carbon vacancy. In Figure 5.44b, the acceptor and donor levels of the carbon vacancy in 4H-SiC that were determined by photo-EPR and those from theoretical calculations are also shown. As shown in Figure 5.44b, the EH7 center has been identified as the donor level of a carbon vacancy, based on these experimental results and theoretical study [160].
  4. Systematic DLTS studies have shown that the E1/E2 c05-math-0532 and R centers c05-math-0533 that are observed in 6H-SiC are equivalent to the c05-math-0534 and EH7 centers in 4H-SiC, respectively [159]. Additionally, the K3 center c05-math-0535 that is observed in 3C-SiC should have the same origin as that of the EH7 center in 4H-SiC [164], that is, a carbon vacancy. Figure 5.45 shows the energy positions of these deep levels in three SiC polytypes. The energy positions for the different charge states of carbon vacancies in the different SiC polytypes are aligned with respect to the valence band edge, and are thus scaled with respect to the conduction band edge according to the bandgap. This is expected, because the vacancy causes a localized disturbance of the valence electrons.
  5. The c05-math-0536 and EH6/7 centers in 4H-SiC are generated through low-energy electron irradiation, by which only the carbon atoms are displaced, and no thermal treatment is required after the irradiation process to form the defect centers [158, 165]. The threshold electron energy for the generation of these levels is around 95–100 keV, which agrees with the threshold energy that was theoretically estimated for carbon-atom displacement (102 keV), as shown in Figure 5.46 [166]. Although several carbon-interstitial-related deep levels are also created by the electron irradiation, these levels are annealed out by thermal treatment at 1000 °C, probably because of the out-diffusion of the carbon interstitials. The densities of both the carbon-vacancy-related and carbon-interstitial-related defects are almost in proportion to the electron fluence, and the defect density can exceed that of any impurities in the SiC epitaxial layers, indicating the exclusion of the involvement of impurities.
  6. In the as-grown epilayers, the densities of the c05-math-0537 and EH6/7 centers increase significantly when the epilayer is grown under Si-rich conditions, but decrease when grown under C-rich conditions [167, 168], which is consistent with the identified defect origin (carbon vacancy).
  7. The c05-math-0538 and EH6/7 centers in 4H-SiC are thermally stable [152, 169], but the densities of the c05-math-0539 and EH6/7 centers in the epitaxial layers are found to increase rapidly through thermal annealing in Ar at temperatures above 1750 °C [170]. Figure 5.47a shows the density of the c05-math-0540 center in several 4H-SiC epitaxial layers annealed at various temperatures. The Arrhenius plot of the density of the c05-math-0541 center as a function of the inverse of the annealing temperature is also shown in Figure 5.47b. Here, the annealing was performed in pure Ar for 30 min. When an epitaxial layer with a very low initial c05-math-0542 center density c05-math-0543 was used, the c05-math-0544 density had already started to increase when the annealing temperature was 1600 °C. When the initial c05-math-0545 density was rather high c05-math-0546, the c05-math-0547 density gradually decreased until the temperature reached 1700 °C, but then increased remarkably after reaching a minimum value. After annealing at temperatures of more than 1750 °C, the c05-math-0548 densities for all samples fall into the same line, as shown in Figure 5.47. The activation energy obtained from the slope is approximately 5.8 eV.
  8. As described in Section 4.3.2, the densities of the c05-math-0549 and EH6/7 centers in the epitaxial layers depend strongly on the growth temperature. Figure 5.48 shows the c05-math-0550 density for as-grown 4H-SiC(0001) epitaxial layers that were grown at various temperatures. The densities of the c05-math-0551 (and EH6/7) centers are reduced with increasing the high C/Si ratio, and the minimum c05-math-0552 density obtained at each growth temperature actually coincides with the data obtained from the annealing experiment (Figure 5.47): the dotted line in Figure 5.48 is the same as that plotted in Figure 5.47a. Therefore, this line, with an activation energy of about 5.8 eV, is universal in SiC, and may correspond to the equilibrium density of the carbon vacancy in SiC.
c05f042

Figure 5.42 (a) DLTS peaks originating from the c05-math-0502 center in n-type 4H-SiC. The “negative-U” property is revealed by using light illumination and short filling pulse ( [156] reproduced with permission from American Physical Society). (b) Energy levels of the c05-math-0503 center in n-type 4H-SiC.

c05f043

Figure 5.43 Densities of the c05-math-0504 center versus those of the EH6/7 center for as-grown, electron-irradiated, and annealed 4H-SiC [158, 159]. The densities of both deep levels are varied by changing the epitaxial growth conditions, the electron energy and fluence during irradiation, and the annealing conditions.

c05f044

Figure 5.44 (a) Area density of a single-negatively-charged carbon vacancy c05-math-0505 determined by EPR versus that of the c05-math-0506 center taken from the depth profile determined by DLTS measurements [163]. (b) Comparison of energy levels of the c05-math-0507 and EH6/7 centers obtained by DLTS and that those of carbon vacancy in 4H-SiC (photo-EPR and theory).

c05f045

Figure 5.45 Energy positions of deep levels originating from a carbon vacancy in 4H-, 6H-, and 3C-SiC.

c05f046

Figure 5.46 Generation rate of the c05-math-0508 and EH6/7 centers versus irradiation energy of electrons ( [166] reproduced with permission fromAIP Publishing LLC).

c05f047

Figure 5.47 (a) Density of the c05-math-0553 center in several 4H-SiC epitaxial layers annealed at various temperatures and (b) Arrhenius plot of the density of the c05-math-0554 center as a function of the inverse of the annealing temperature ( [170] reproduced with permission from AIP Publishing LLC).

c05f048

Figure 5.48 c05-math-0555 density for as-grown 4H-SiC(0001) epitaxial layers grown at various temperatures.

The c05-math-0556 center is the most important defect, because of its abundance and because of its property of being a major carrier-lifetime killer [171, 172]. To obtain a long carrier lifetime, which is beneficial for reduction of the on-resistance in bipolar devices, the density of the c05-math-0557 center must be decreased to the order of c05-math-0558 or lower. Thus far, two successful techniques have been proposed to eliminate the c05-math-0559 center.

In the first technique, excess carbon atoms are introduced from the outside and the diffusion of these carbon atoms into the bulk region is promoted by high-temperature annealing [173, 174]. This can actually be achieved by carbon ion-implantation and subsequent annealing in Ar at 1650–1700 °C. Figure 5.49 shows the DLTS spectra obtained from n-type 4H-SiC epitaxial layers before and after carbon ion-implantation followed by high-temperature Ar annealing. Both the c05-math-0560 and EH6/7 centers are eliminated by this process down to the detection limit (approximately c05-math-0561 in this case). Because the created carbon interstitials have large diffusion coefficients above 1400–1500 °C, the excess carbon interstitials diffuse and fill the carbon vacancies, leading to the elimination of the c05-math-0562 and EH6/7 centers from the surface region to the bulk. After annealing, the defective implanted region near the surface is removed by plasma etching.

c05f049

Figure 5.49 DLTS spectra obtained from n-type 4H-SiC epitaxial layers before and after carbon ion-implantation followed by high-temperature Ar annealing [173].

In the second technique, thermal oxidation of SiC under appropriate conditions results in the elimination of the c05-math-0563 and EH6/7 centers. During thermal oxidation, the carbon atoms are mostly removed by the creation of carbon monoxide (CO). However, some portion of the carbon atoms is emitted to the bulk region and diffuses into a deep region of SiC. Because silicon atoms are emitted during thermal oxidation of Si [175], it is expected that carbon (and also silicon) atoms will be emitted during the thermal oxidation of SiC. As in the case of the carbon ion implantation, the diffusion coefficient of the carbon interstitials is so large that the depth to which the c05-math-0564 and EH6/7 centers are eliminated can exceed c05-math-0565, by either the combination of oxidation and subsequent high-temperature Ar annealing or by high-temperature oxidation (1300–1400 °C) [176–178]. Note that the thermal oxide is removed before Ar annealing. It was also reported that the migration energy of the carbon interstitials is much smaller than that of the silicon interstitials [179], and the self-diffusion coefficient of carbon is much larger than that of silicon [180]. Figure 5.50 shows the depth profile of the c05-math-0566 density for the 4H-SiC epitaxial layers after various thermal treatments (e.g., oxidation, Ar annealing) [178]. By extending the oxidation time or raising the oxidation temperature, the “c05-math-0567” region extending from the surface becomes thicker (c05-math-0568 or even thicker).

c05f050

Figure 5.50 (a,b) Depth profile of the c05-math-0569 density for the 4H-SiC epitaxial layers after various thermal treatments (e.g., oxidation, Ar annealing) ( [178] reproduced with permission from AIP Publishing LLC).

In both the deep-level reduction techniques (carbon ion implantation and thermal oxidation), the depth of the “c05-math-0570” region depends not only on the process conditions but also on the initial c05-math-0571 density. When the initial c05-math-0572 density is relatively low c05-math-0573, it is easy to eliminate these deep levels in c05-math-0574 regions. However, a thick “c05-math-0575” region can hardly be obtained if the initial c05-math-0576 density is high c05-math-0577. The depth profiles of the c05-math-0578 density after these defect-reduction processes can be predicted through numerical simulation of the carbon diffusion process [178]. It was noted that carbon diffusion into the SiC bulk region results in the generation of a few new deep levels, such as the ON1 c05-math-0579, ON2 c05-math-0580, and HK0 c05-math-0581 centers [181, 182], which are commonly observed near the surface in both carbon-implanted SiC and oxidized SiC. The HK0 center is annealed out by thermal treatment at temperatures above 1350–1400 °C, while the ON1/ON2 centers (the same defect, but with different charge states) survive even after annealing at 1800 °C. A carbon-interstitial-related defect such as the carbon di-interstitial c05-math-0582 was suggested as the origin of the HK0 center [182].

For details of other intrinsic deep levels in SiC, such as c05-math-0583 [152], UT1 c05-math-0584 [183], HS2 c05-math-0585 [165], and HK4 c05-math-0586 [154], please see the individual references provided or review papers [152, 184]. The exact correlation between the observed deep levels and a specific defect structure (e.g., silicon vacancy, antisite) has not yet been established. Figure 5.51 shows the energy positions of the various intrinsic defects that are expected in 4H-SiC, which were obtained through theoretical calculations [179, 185, 186]. The energy levels of the silicon vacancy c05-math-0587, the carbon antisite–carbon vacancy pair c05-math-0588, the carbon di-interstitial c05-math-0589, and other defects have been reported. The diffusion of these intrinsic defects has also been investigated through calculations [179]. However, the experimental assignment of actual deep levels to these defects has not yet been achieved.

c05f051

Figure 5.51 Energy positions of the various intrinsic defects that are expected in 4H-SiC, which were obtained through theoretical calculations [179, 185, 186].

5.3.1.2 Impurities

The purity of the SiC epitaxial layers grown by chemical vapor deposition (CVD) is rather high, although the substrates do contain various impurities, including heavy metals. Other than the intentional dopants, the major impurities that are present in SiC epitaxial layers are boron and titanium, as described in Section 4.3.2. The typical densities of these impurities are c05-math-0590 for boron and c05-math-0591 for titanium. Several other impurities are also occasionally observed. Figure 5.52 shows the energy positions of the major metallic impurity-related deep levels that are observed in 4H-SiC.

c05f052

Figure 5.52 Energy positions of the major metallic-impurity-related deep levels experimentally observed in 4H-SiC.

Boron (B) atoms substituting the silicon lattice sites create boron acceptors c05-math-0592 [187]. However, some boron atoms form pairs with carbon vacancies c05-math-0593 [188], leading to the formation of a “deep boron” level; this is the D center in 4H-SiC c05-math-0594 [187]. The energy position of the D center is deeper in 6H-SiC c05-math-0595. The D center is also acceptor-like [189], and contributes to the DAP luminescence at relatively long wavelengths. The generation of the D center in SiC epitaxial layers is enhanced under Si-rich growth conditions and is reduced under C-rich conditions [20]. The D center is thermally stable, but can be reduced by thermal oxidation or by carbon ion implantation, which is reasonable because interaction of the D center c05-math-0596 with a carbon interstitial will naturally annihilate the defect.

Oxygen (O) is a common impurity in almost all semiconductor materials, including Si and GaAs. Because nitrogen is a dominant residual impurity in high-purity SiC, oxygen contamination is naturally expected. However, it was found that oxygen is hardly incorporated in SiC during epitaxial growth, even when an oxygen-containing gas is intentionally introduced into the CVD reactor [190, 191]. Etching of c05-math-0597 or c05-math-0598 by hydrogen that is supplied as a carrier gas, or decomposition of c05-math-0599 at the high growth temperatures may suppress oxygen incorporation. No deep levels that could be attributed to pure oxygen-related defects have been identified. The oxygen density in SiC seems to be very low (c05-math-0600 or lower).

Titanium (Ti) is another common impurity in SiC, and it creates very shallow electron traps near the conduction band edge c05-math-0601 in 4H-SiC [29]. Figure 5.53 shows schematically the energy positions of the Ti levels in the different SiC polytypes. The Ti levels with respect to EV are almost identical when plotted in the bandgaps of the different polytypes, indicating that the Langer–Heinrich rule [192] is valid for Ti in SiC. Ti is electrically active in 4H-SiC, but its acceptor levels are resonant in the conduction band for 6H-, 15R-, and 3C-SiC. As described in Section 5.1.1, Ti exhibits a unique greenish luminescence in 4H-, 6H-, and 15R-SiC, but not in 3C-SiC.

c05f053

Figure 5.53 Energy positions of the Ti levels in the different SiC polytypes ( [29] reproduced with permission from American Physical Society).

Vanadium (V) is an amphoteric impurity, and it forms an acceptor level in the upper half of the bandgap and a donor level in the lower half of the bandgap. In 4H-SiC, the acceptor level and the donor level are located at c05-math-0602 and c05-math-0603, respectively [31]. Thus, vanadium creates a very deep level, especially in p-type SiC, and is thermally stable. This is why vanadium doping can be used to form semi-insulating SiC, as described in Section 3.4.4. In SiC epitaxial layers grown by CVD, however, the vanadium density is usually less than c05-math-0604.

The deep levels of chromium (Cr), molybdenum (Mo), tungsten (W), and iron (Fe) in 4H-SiC have been identified [7, 193–195]. In infrared PL spectra from 4H-SiC, the sharp PL peaks at 1170 and 1171 nm (1.0586 and 1.0595 eV) are well known and have been assigned to a UD center (UD1). In recent years, systematic investigations of this defect center by PL, EPR, DLTS, and theoretical calculations have revealed that the origin of UD1 is actually tungsten [36]. In high-purity SiC epitaxial layers, however, the densities of Cr, Mo, W, and Fe are lower than c05-math-0605. It was demonstrated that some transition metals such as Ni and Cr exhibit significant diffusion in SiC at temperatures higher than 1500 °C [196].

5.3.2 Carrier Lifetime Killer

Correlation between the carrier lifetimes and the density of the c05-math-0606 center was suggested at an early stage of the studies of SiC [167, 197] and, in recent years, the c05-math-0607 center has been unambiguously identified as a major carrier lifetime killer, at least, in n-type 4H-SiC [171, 172]. Figure 5.54 shows the inverse of the carrier lifetime versus the measured density of the c05-math-0608 center for c05-math-0609 n-type 4H-SiC epitaxial layers [172, 198]. When the c05-math-0610 density is higher than c05-math-0611, the inverse of the carrier lifetime is proportional to the c05-math-0612 density, indicating that the lifetime is governed by Shockley–Read–Hall (SRH) recombination via the c05-math-0613 center. However, the correlation between the lifetime and the c05-math-0614 density is unclear when the c05-math-0615 density is in the c05-math-0616 range. As described in Section 5.1.4, multiple recombination processes exist, including SRH recombination and several other recombination processes. Thus, the carrier lifetime c05-math-0617 can be expressed by the following equation:

where c05-math-0619 is the SRH lifetime governed by the recombination centers, and c05-math-0620 is the carrier lifetime that is governed by the other recombination processes, such as surface recombination, recombination in the substrate, Auger recombination, and recombination at extended defects. Here, the inverse of c05-math-0621 is proportional to the density of the recombination centers (c05-math-0622, where c05-math-0623 is a constant, and c05-math-0624 is the density of the c05-math-0625 center), while c05-math-0626 can be assumed to be independent of the c05-math-0627 density. By using the model expressed by Equation 5.19, the experimental data can be fitted. The fitted result is shown in Figure 5.54 as a solid line for c05-math-0628, and as two broken lines for c05-math-0629 and c05-math-0630. Here the c05-math-0631 line is determined by recombination at the surface and in the substrate, due to the limited thickness of epitaxial layers (c05-math-0632 in this plot), and this component decreases by using thicker epitaxial layers. The detail was explained in Section 5.1.4.4.

c05f054

Figure 5.54 Inverse of the carrier lifetime versus the measured density of the c05-math-0633 center for c05-math-0634 n-type 4H-SiC epitaxial layers [172, 198].

Based on the results described above, a simple relationship between the high-injection bulk lifetime c05-math-0635 and the c05-math-0636 density c05-math-0637 can be established as follows:

5.20 equation

Note that this equation is valid when Auger recombination and the recombination at extended defects are less important, and it is independent of both the epilayer thickness and the SRV (surface recombination velocity). Instead, the factor of c05-math-0639 is a function of the excitation intensity (i.e., the injection level) and the temperature.

In semiconductor textbooks, it is stated that a midgap level is an effective recombination center and can thus be a lifetime killer [157]. In 4H-SiC, however, the midgap level of the EH6/7 c05-math-0640 center does not work as a carrier lifetime killer [199]. This can be deduced by accounting for the properties of the c05-math-0641 and EH6/7 (EH7) centers. As described in Section 5.3.1, the origin of both the c05-math-0642 and EH7 centers is a carbon vacancy c05-math-0643 with different charge states:

  1. c05-math-0644, negative c05-math-0645 (trap and emit two electrons), c05-math-0646
  2. c05-math-0647.

In n-type 4H-SiC, the Fermi level is well above the c05-math-0648 center, and the c05-math-0649 defect is charged to c05-math-0650. When excess electron–hole pairs are generated, a hole is first captured at the c05-math-0651 defect, because of the large hole capture cross-section (the attractive Coulomb force between c05-math-0652 and the hole). Then, the c05-math-0653 defect changes its charge state as follows: c05-math-0654. If the hole density is extremely high, then c05-math-0655 may proceed. Under these circumstances, the capture cross-sections of the electrons for c05-math-0656 and c05-math-0657 are very large, because of its c05-math-0658 property. Thus, the c05-math-0659 defect immediately tends to revert to the original state; c05-math-0660 or c05-math-0661, and subsequently c05-math-0662. These capture and recombination processes are repeated upon electron–hole excitation. Therefore, only the c05-math-0663 center c05-math-0664 is involved in the recombination process. The EH7 center c05-math-0665 does not appear during this process. This is why the EH7 center is not important in carrier recombination in n-type 4H-SiC. The EH7 center may, however, be important in p-type 4H-SiC, where the Fermi level is close to the valence band and the c05-math-0666 defect is charged to c05-math-0667 .

Figure 5.55 shows the c05-math-0671 decay curves at room temperature that were obtained from a c05-math-0672 n-type 4H-SiC epitaxial layer [100]. The decay curves for the as-grown material and that after c05-math-0673 reduction through thermal oxidation at 1400 °C for 48 h are shown. For the as-grown epitaxial layer, the measured lifetime is c05-math-0674, while the lifetime has improved remarkably to c05-math-0675 after the defect reduction process. The depth of the c05-math-0676 region of this particular sample is estimated to be about c05-math-0677, indicating that the defect is eliminated throughout the entire thickness of the epitaxial layer. After lifetime measurement of the defect-eliminated sample, the surface was passivated with a 20-nm-thick deposited oxide and annealed in nitric oxide (NO) at 1250 °C for 30 min. After this surface passivation, the lifetime increased to c05-math-0678, and the lifetime increased further to c05-math-0679 at a measurement temperature of 200 °C. Because a relatively low interface state density is obtained over the whole energy range of the bandgap by this passivation process [200], surface recombination may be suppressed, leading to a longer carrier lifetime. Similar carrier lifetimes can be achieved by carbon ion implantation and subsequent Ar annealing [201].

c05f055

Figure 5.55 c05-math-0668 decay curves at room temperature obtained from a c05-math-0669 n-type 4H-SiC epitaxial layer ( [100] reproduced with permission from The Japan Society of Applied Physics). The decay curves for the as-grown material and that after c05-math-0670-center reduction through thermal oxidation at 1400 °C for 48 h are shown.

In general, the experimentally determined carrier lifetimes are underestimated, because of parasitic recombination paths, as described in Section 5.1.4. In Figure 5.56, the measured lifetimes are plotted as a function of the epitaxial layer thickness. The results for samples taken before oxidation (as-grown), after oxidation, and with the subsequent surface passivation are indicated by closed circles, open circles, and closed triangles, respectively. In the oxidized samples, sufficiently long-term oxidation was performed at 1300–1400 °C to ensure that the c05-math-0680 center is eliminated throughout the entire thickness of the epitaxial layers. The lifetimes of the oxidized samples increase rapidly with increasing epitaxial layer thickness, while the thickness dependence is very small for the as-grown samples. This result indicates that the measured lifetime is severely affected by the carrier recombination in the substrate (or near the epitaxial layer/substrate interface) when the bulk lifetime becomes very long because of c05-math-0681 elimination, which is consistent with Figure 5.17. In Figure 5.56, the simulated dependence of the lifetime on the epilayer thickness is also indicated by dashed lines for the various bulk lifetimes of the epitaxial layers. The SRV was assumed to be c05-math-0682. As determined from Figure 5.56, the real bulk lifetime c05-math-0683 will be longer than c05-math-0684. The ambipolar diffusion length c05-math-0685 is given by the following equation [1, 151]:

5.21 equation
5.22 equation

Here, c05-math-0688 and c05-math-0689 are the diffusion coefficients of the electrons and the holes, respectively. Ambipolar diffusion is described in greater detail in Section 7.3.1. By accounting for the diffusion coefficients and assuming a long carrier lifetime of c05-math-0690, the diffusion length of the carriers is estimated to be longer than c05-math-0691 in the c05-math-0692 region.

c05f056

Figure 5.56 Measured lifetimes versus epitaxial layer thickness. The results for samples taken before oxidation (as-grown), after oxidation, and with the subsequent surface passivation are indicated by closed circles, open circles, and closed triangles, respectively.

In Si, heavy metallic impurities, such as Au, Pt, and Fe, create midgap levels that act as efficient recombination centers [151]. Although the deep levels of several metallic impurities (V, Cr, W) are known in SiC, the densities of these impurity levels determined by DLTS measurements are usually close to the detection limit (about c05-math-0693). Therefore, it can be concluded that the SRH recombination center (i.e., the lifetime killer) is indeed the c05-math-0694 center in n-type 4H-SiC.

The carrier lifetime in p-type SiC is more complicated. Because the Fermi level is close to the valence band in p-type SiC, the carbon vacancy defect must be positively charged (c05-math-0695: EH7 center) in equilibrium. Thus, it is expected that any excess electrons are quickly trapped by the defect, and the charge state will change to neutral. However, accurate capture cross-sections for the electrons and holes are not yet known for this neutral defect. Figure 5.57 shows the c05-math-0696 decay curves at room temperature that were obtained from a c05-math-0697, lightly-doped p-type 4H-SiC epitaxial layer [202, 203]. The decay curves for the as-grown material and for the material after carbon-vacancy reduction via thermal oxidation at 1400 °C for 48 h are shown. The measured lifetime of the as-grown epilayer is c05-math-0698, and the lifetime increased to c05-math-0699 after the defect reduction process. The lifetime improved slightly to c05-math-0700 following surface passivation with a nitrided oxide [202, 203]. Carbon ion implantation followed by Ar annealing at 1650 °C, instead of thermal oxidation, provides a very similar result. Thus, reduction of the carbon vacancies (i.e., the c05-math-0701 and EH6/7 centers) is effective for improving the carrier lifetimes in p-type SiC, but the level of improvement is much smaller than that in n-type SiC. Further investigations are required to identify the lifetime killer and enhance the carrier lifetime in p-type SiC.

c05f057

Figure 5.57 c05-math-0702 decay curves at room temperature obtained from a c05-math-0703, lightly-doped p-type 4H-SiC epitaxial layer [202, 203]. The decay curves for the as-grown material and for the material after carbon-vacancy reduction via thermal oxidation at 1400 °C for 48 h are shown.

The effects of extended defects on carrier lifetimes have also been investigated [68, 204, 205]. All the extended defects cause local reduction of the carrier lifetimes near the defects. This is already indicated in the PL mapping data shown in Figure 5.26, where the PL intensities of the free exciton peaks are greatly reduced near the dislocations. The magnitude of the reduction in PL intensity is much larger near TSDs than TEDs. Figure 5.58 shows the PL decay curves that were measured from a c05-math-0704 n-type 4H-SiC(0001) epitaxial layer at room temperature. The measurements were conducted at a defect-free region, near a TSD core, a TED core, and inside a single SSF expanded from a BPD. The carrier lifetime obtained from the PL decay was c05-math-0705 from the defect-free region, c05-math-0706 near the TED, c05-math-0707 near the TSD, and c05-math-0708 for the SSF. Thus, the stacking faults (nucleated from BPDs) and the TSDs are the most detrimental defects for the carrier lifetimes in SiC. The effects of these extended defects depend strongly on the crystal quality, and are relatively larger in high-quality SiC with long carrier lifetimes.

c05f058

Figure 5.58 PL decay curves that were measured from a c05-math-0709 n-type 4H-SiC(0001) epitaxial layer at room temperature. The measurements were conducted at a defect-free region, near a TSD core, a TED core, and inside a single SSF expanded from a BPD.

5.3.2.1 Lifetime Control

Because the density of the c05-math-0710 (and the EH6/7) center can be intentionally increased by low-energy electron irradiation [158, 165], lifetime control is easily achieved in n-type SiC. Although several deep levels, which may be ascribed to the interstitial-related defects, are also generated by electron irradiation, any deep levels other than the c05-math-0711 and EH6/7 centers are eliminated by annealing at 900–1000 °C [159, 165]. Figure 5.59 shows the density of the c05-math-0712 center generated by electron irradiation (after 950 °C annealing) versus the electron fluence [158, 163]. The electron irradiation was carried out at energies of 116, 160, 200, or 250 keV. In either case, the c05-math-0713 center density increases almost in proportion to the electron fluence. By increasing the fluence from c05-math-0714 to c05-math-0715 along with the irradiation energy, the c05-math-0716 center density can be controlled over a very wide range from approximately c05-math-0717 to c05-math-0718.

c05f059

Figure 5.59 Density of the c05-math-0719 center generated by electron irradiation (after 950 °C annealing) versus the electron fluence [158, 163]. The electron irradiation was carried out at energies of 116, 160, 200, or 250 keV.

Figure 5.60 shows a map of the carrier lifetimes for n-type 4H-SiC epitaxial layers after electron irradiation and subsequent annealing in Ar at 950 °C for 30 min. The starting material was a c05-math-0720 4H-SiC epitaxial layer, grown on an n-type substrate and doped to c05-math-0721. Electron irradiation was performed at 160 keV with the fluence range from c05-math-0722 to c05-math-0723 on the sample being controlled by using a copper mask during irradiation. Because selective electron irradiation was performed, different c05-math-0724 densities ranging from c05-math-0725 to c05-math-0726 over six different areas were realized in the sample. The carrier lifetime is shorter in the area where the electron fluence is higher (i.e., the c05-math-0727 density is higher). More importantly, a highly uniform carrier lifetime can be realized in each area. These results indicate that the carrier lifetime can be controlled reasonably well by the electron irradiation process. This technique is useful for obtaining very uniform carrier lifetimes or for the reduction of the switching losses in SiC bipolar devices. Control of the carrier lifetimes by similar low-energy electron irradiation is also possible for p-type 4H-SiC [203], suggesting that a carbon vacancy (e.g., the c05-math-0728 or EH6/7 centers) can be a lifetime killer when its density is sufficiently high c05-math-0729.

c05f060

Figure 5.60 (a) Electron fluence and resulting c05-math-0730 center density in a c05-math-0731 n-type 4H-SiC epitaxial layer doped to c05-math-0732 and (b) map of the carrier lifetimes for the selectively irradiated 4H-SiC epitaxial layers after electron irradiation and subsequent annealing in Ar at 950 °C for 30 min.

5.4 Summary

Basic characterization techniques have been described in this chapter with emphasis placed on the special care that is required for the characterization of SiC. PL is a powerful technique for characterization of optical properties, and can reflect the incorporation of specific impurities (including dopants), point defects, and any other localized levels. Raman scattering is useful for the identification of the SiC polytypes and for characterization of strains. c05-math-0733 measurements give the net doping density c05-math-0734, while the carrier density c05-math-0735 can be measured directly only by Hall effect measurements. The carrier lifetimes can be determined by several techniques, but the decay time obtained does not always mean the carrier lifetime. Special attention must also be paid to the effects of surface recombination and recombination in the underlying substrate.

Extended defects have been detected via destructive methods such as molten KOH etching. High-resolution X-ray topography and PL mapping/imaging serve as attractive techniques for nondestructive identification of the locations and types of extended defects. Deep levels, which originate from point defects or impurities, can be characterized by DLTS, and high-temperature (up to 800 K) measurements are required to monitor the midgap levels in SiC. To identify the origin of a deep level, a detailed comparison study of the results of DLTS, EPR measurements, and theoretical calculations is required.

Because micropipe defects have been eliminated, the macroscopic defects that are generated during epitaxial growth, such as triangular defects, carrot defects, in-grown stacking faults, and particles, are the most detrimental defects for any SiC device. A BPD is split into two partial dislocations with a single SSF in between them. This stacking fault is then expanded upon carrier injection and recombination, leading to a considerable reduction of the carrier lifetimes and increased leakage current. Therefore, BPDs cause the degradation of bipolar-type SiC devices (but not unipolar devices). The effects of TSDs on the breakdown voltage and leakage current are very small when no surface pits are formed at the TSD locations. The effects of TEDs are mostly negligibly small. The major carrier lifetime killer has been identified as the c05-math-0736 center c05-math-0737, which is the acceptor level of a carbon vacancy, in n-type 4H-SiC. In p-type 4H-SiC, however, the carrier-lifetime killers have not yet been conclusively identified. TSDs and TEDs cause local reduction of the carrier lifetimes, while BPDs induce expansion of the SSFs upon carrier injection, leading to significant reduction of the carrier lifetimes in the fault region. However, further investigations are required to fully clarify and understand the behavior of the extended and point defects in SiC.

The present chapter did not describe the characterization of SiC by other techniques, such as standard X-ray diffraction, TEM, secondary ion mass spectrometry (SIMS), X-ray photoelectron spectroscopy (XPS), and Auger electron spectroscopy (AES). These techniques are frequently used for material characterization but special care and unique analyses are not always required for SiC. Please see individual textbooks for details of these techniques.

References

  1. [1] Schroder, D.K. (2006) Semiconductor Material and Device Characterization, 3rd edn, Wiley-IEEE.
  2. [2] Perkowitz, S. (1993) Optical Characterization of Semiconductors, Academic Press.
  3. [3] Choyke, W.J. (1969) Optical properties of polytypes of SiC: Interband absorption, and luminescence of nitrogen-exciton complexes. Mater. Res. Bull., 4, 141.
  4. [4] Devaty, R.P. and Choyke, W.J. (1997) Optical characterization of silicon carbide polytypes. Phys. Status Solidi A, 162, 5.
  5. [5] Choyke, W.J. and Devaty, R.P. (2004) Optical properties of SiC: 1997–2002, in Silicon Carbide—Recent Major Advances (eds W.J. Choyke, H. Matsunami and G. Pensl), Springer, p. 413.
  6. [6] Egilsson, T., Ivanov, I.G., Son, N.T. et al. (2004) Exciton and defect photoluminescence from SiC, in Silicon Carbide, Materials, Processing, and Devices (eds Z.C. Feng and J.H. Zhao), Taylor & Francis Group, p. 81.
  7. [7] Janzen, E., Gali, A., Henry, A. et al. (2008) Defects in SiC, in Defects in Microelectronic Materials and Devices, Taylor & Francis Group, p. 615.
  8. [8] Ivanov, I.G., Zhang, J., Storasta, L. and Janzen, E. (2002) Photoconductivity of lightly-doped and semi-insulating 4H-SiC and the free exciton binding energy. Mater. Sci. Forum, 389–393, 613.
  9. [9] Ivanov, I.G., Hallin, C., Henry, A. et al. (1996) Nitrogen doping concentration as determined by photoluminescence in 4H- and 6H-SiC. J. Appl. Phys., 80, 3504.
  10. [10] Henry, A., Forsberg, U., Linnarsson, M.K. and Janzen, E. (2005) Determination of nitrogen doping concentration in doped 4H-SiC epilayers by low temperature photoluminescence. Phys. Scr., 72, 254.
  11. [11] Bebb, H.B. and Williams, E.W. (1972) Photoluminescence, in Semiconductors and Semimetals, vol. 8 (eds R.K. Willardson and A.C. Beer), Academic Press.
  12. [12] Itoh, A., Kimoto, T. and Matsunami, H. (1996) Exciton-related photoluminescence in 4H-SiC grown by step-controlled epitaxy. Jpn. J. Appl. Phys., 35, 4373.
  13. [13] Choyke, W.J., Hamilton, D.R. and Patrick, L. (1964) Optical properties of cubic SiC: luminescence of nitrogen-exciton complexes, and interband absorption. Phys. Rev., 133, A1163.
  14. [14] Patrick, L., Choyke, W.J. and Hamilton, D.R. (1965) Luminescence of 4H SiC, and location of conduction-band minima in SiC polytypes. Phys. Rev., 137, A1515.
  15. [15] Choyke, W.J. and Patrick, L. (1962) Exciton recombination radiation and phonon spectrum of 6H SiC. Phys. Rev., 127, 1868.
  16. [16] Kimoto, T., Nakazawa, S., Hashimoto, K. and Matsunami, H. (2001) Reduction of doping and trap concentrations in 4H-SiC epitaxial layers grown by chemical vapor deposition. Appl. Phys. Lett., 79, 2761.
  17. [17] Henry, A. and Janzen, E. (2006) Photoluminescence of phosphorous doped SiC. Mater. Sci. Forum, 527–529, 589.
  18. [18] Clemen, L.L., Devaty, R.P., MacMillan, M.F. et al. (1993) Aluminum acceptor four particle bound exciton complex in 4H, 6H, and 3C SiC. Appl. Phys. Lett., 62, 2953.
  19. [19] Pedersen, H., Henry, A., Hassan, J. et al. (2007) Growth and photoluminescence study of aluminium doped SiC epitaxial layers. Mater. Sci. Forum, 556–557, 97.
  20. [20] Sridhara, S.G., Clemen, L.L., Devaty, R.P. et al. (1998) Photoluminescence and transport studies of boron in 4H SiC. J. Appl. Phys., 83, 7909.
  21. [21] Suzuki, A., Matsunami, H. and Tanaka, T. (1977) Photoluminescence spectra of Ga-doped and Al-doped 4H-SiC. J. Phys. Chem. Solids, 38, 693.
  22. [22] Choyke, W.J. and Patrick, L. (1970) Luminescence of donor-acceptor pairs in cubic SiC. Phys. Rev. B, 2, 4959.
  23. [23] Choyke, W.J. (1990) Optical and electronic properties of SiC, in The Physics and Chemistry of Carbides, Nitrides, and Borides (ed R. Freer), Kluwer Academic Publishers, p. 563.
  24. [24] Ikeda, M., Matsunami, H. and Tanaka, T. (1980) Site effect on the impurity levels in 4H, 6H, and 15R SiC. Phys. Rev. B, 22, 2842.
  25. [25] Yamada, S. and Kuwabara, S. (1974) Photoluminescence of β-SiC doped with boron and nitrogen, in Silicon Carbide 1973 (eds R.C. Marshall, J.W. Faust Jr. and C.E. Ryan), University of South Carolina Press, p. 305.
  26. [26] Ivanov, I.G., Henry, A. and Janzen, E. (2005) Ionization energies of phosphorus and nitrogen donors and aluminum acceptors in 4H silicon carbide from the donor-acceptor pair emission. Phys. Rev. B, 71, 241201(R).
  27. [27] van Kemenade, A.W.C. and Hagen, S.H. (1974) Proof of the involvement of Ti in the low-temperature abc luminescence spectrum of 6H SiC. Solid State Commun., 14, 1331.
  28. [28] Lee, K.M., Dang, L.S., Watkins, G.D. and Choyke, W.J. (1985) Optically detected magnetic resonance study of SiC: Ti. Phys. Rev. B, 32, 2273.
  29. [29] Dalibor, T., Pensl, G., Nordell, N. and Schöner, A. (1997) Electrical properties of the titanium acceptor in silicon carbide. Phys. Rev. B, 55, 13618.
  30. [30] Henry, A. and Janzen, E. (2006) Titanium related luminescence in SiC. Superlattices Microstruct., 40, 328.
  31. [31] Schneider, J., Muller, H.D., Maier, K. and Fuchs, F. (1990) Infrared spectra and electron spin resonance of vanadium deep level impurities in silicon carbide. Appl. Phys. Lett., 56, 1184.
  32. [32] Lauer, V., Bremond, G., Souifi, A. et al. (1999) Electrical and optical characterisation of vanadium in 4H and 6H-SiC. Mater. Sci. Eng., B, 61–62, 248.
  33. [33] Prezzi, D., Eberlein, T.A.G., Filhol, J.-S. et al. (2004) Optical and electrical properties of vanadium and erbium in 4H-SiC. Phys. Rev. B, 69, 193202.
  34. [34] Kunzer, M., Dombrowski, K.F., Fuchs, F. et al. (1996) Identification of optically and electrically active molybdenum trace impurities in 6H-SiC substrates. Inst. Phys. Conf. Ser., 142, 385.
  35. [35] Son, N.T. (1999) Photoluminescence and Zeeman effect in chromium-doped 4H and 6H SiC. J. Appl. Phys., 86, 4348.
  36. [36] Gällström, A., Magnusson, B., Beyer, F. et al. (2012) Optical identification and electronic configuration of tungsten in 4H- and 6H-SiC. Physica B, 407, 1462.
  37. [37] Magnusson, B. and Janzen, E. (2005) Optical characterization of deep level defects in SiC. Mater. Sci. Forum, 483–485, 341.
  38. [38] Wagner, M., Magnusson, B., Chen, W.M. et al. (2000) Electronic structure of the neutral silicon vacancy in 4H and 6H SiC. Phys. Rev. B, 62, 16555.
  39. [39] Janzen, E., Ivanov, I.G., Son, N.T. et al. (2003) Defects in SiC. Physica B, 340–342, 15.
  40. [40] Son, N.T., Carlsson, P., ul Hassan, J. et al. (2007) Defects and carrier compensation in semi-insulating 4H-SiC substrates. Phys. Rev. B, 75, 155204.
  41. [41] Egilsson, T., Henry, A., Ivanov, I.G. et al. (1999) Photoluminescence of electron-irradiated 4H-SiC. Phys. Rev. B, 59, 8008.
  42. [42] Steeds, J.W., Evans, G.A., Danks, L.R. et al. (2002) Transmission electron microscope radiation damage of 4H and 6H SiC studied by photoluminescence spectroscopy. Diamond Relat. Mater., 11, 1923.
  43. [43] Steeds, J.W. and Sullivan, W. (2008) Identification of antisite carbon split-interstitial defects in 4H-SiC. Phys. Rev. B, 77, 195204.
  44. [44] Mattausch, A., Bockstedte, M., Pankratov, O. et al. (2006) Thermally stable carbon-related centers in 6H-SiC: Photoluminescence spectra and microscopic models. Phys. Rev. B, 73, 161201(R).
  45. [45] Patrick, L. and Choyke, W.J. (1972) Photoluminescence of radiation defects in ion-implanted 6H SiC. Phys. Rev. B, 5, 3253.
  46. [46] Haberstroh, C., Helbig, R. and Stein, R.A. (1994) Some new features of the photoluminescence of SiC(6H), SiC(4H), and SiC(15R). J. Appl. Phys., 76, 509.
  47. [47] Egilsson, T., Bergman, J.P., Ivanov, I.G. et al. (1999) Properties of the D1 bound exciton in 4H-SiC. Phys. Rev. B, 59, 1956.
  48. [48] Hori, T., Danno, K. and Kimoto, T. (2007) Fast homoepitaxial growth of 4H-SiC with low basal-plane dislocation density and low trap concentration by hot-wall chemical vapor deposition. J. Cryst. Growth, 306, 297.
  49. [49] T. Dalibor, C. Peppermuller, G. Pensl, et al., Inst. Phys. Conf. Ser. 142 1996), 517 Defect centers in ion-implanted 4H silicon carbide.
  50. [50] Storasta, L., Carlsson, F., Sridhara, S.G. et al. (2001) Pseudodonor nature of the DI defect in 4H-SiC. Appl. Phys. Lett., 78, 46.
  51. [51] Gali, A., Deak, P., Rauls, E. et al. (2003) Correlation between the antisite pair and the DI center in SiC. Phys. Rev. B, 67, 155203.
  52. [52] Eberlein, T.A.G., Jones, R., Öberg, S. and Briddon, P.R. (2006) Density functional theory calculation of the DI optical center in SiC. Phys. Rev. B, 74, 144106.
  53. [53] Patrick, L. and Choyke, W.J. (1973) Localized vibrational modes of a persistent defect in ion-implanted SiC. J. Phys. Chem. Solids, 34, 565.
  54. [54] Sridhara, S.G., Nizhner, D.G., Devaty, R.P. et al. (1998) DII revisited in a modern guise—6H and 4H SiC. Mater. Sci. Forum, 264–268, 493.
  55. [55] Mattausch, A., Bockstedte, M. and Pankratov, O. (2004) Carbon antisite clusters in SiC: A possible pathway to the DII center. Phys. Rev. B, 69, 045322.
  56. [56] Feldman, D.W., Parker, J.H. Jr., Choyke, W.J. and Patrick, L. (1968) Phonon dispersion curves by Raman scattering in SiC, polytypes 3C, 4H, 6H, 15R, and 21R. Phys. Rev., 170, 698.
  57. [57] Dubrovskii, G.B. and Lepneva, A.A. (1983) Energy band structure and optical spectra of silicon carbide crystals. Sov. Phys. Solid State, 25, 1330.
  58. [58] Nakashima, S. and Harima, H. (1997) Raman investigation of SiC polytypes. Phys. Status Solidi A, 162, 39.
  59. [59] Nakashima, S. and Harima, H. (2004) Characterization of defects in SiC crystals by Raman scattering, in Silicon Carbide—Recent Major Advances (eds W.J. Choyke, H. Matsunami and G. Pensl), Springer, p. 585.
  60. [60] Nakashima, S. (2004) Raman imaging of semiconductor materials: characterization of static and dynamic properties. J. Phys. Condens. Matter, 16, S25.
  61. [61] Nakashima, S., Okumura, H., Yamamoto, T. and Shimidzu, R. (2004) Deep-ultraviolet Raman microspectroscopy: characterization of wide-gap semiconductors. Appl. Spectrosc., 58, 224.
  62. [62] Harima, H., Nakashima, S. and Uemura, T. (1995) Raman scattering from anisotropic LO-phonon-plasmon–coupled mode in n-type 4H- and 6H-SiC. J. Appl. Phys., 78, 1996.
  63. [63] Lockwood, D.J. and Baribeau, J.-M. (1992) Strain-shift coefficients for phonons in Si1-xGex epilayers on silicon. Phys. Rev. B, 45, 8565.
  64. [64] Hoffmann, H.J. (1979) Defect-level analysis of semiconductors by a new differential evaluation of n (1/T)-characteristics. Appl. Phys., 19, 307.
  65. [65] Matsuura, H., Komeda, M., Kagamihara, S. et al. (2004) Dependence of acceptor levels and hole mobility on acceptor density and temperature in Al-doped p-type 4H-SiC epilayers. J. Appl. Phys., 96, 2708.
  66. [66] Schmidt, F., Krieger, M., Laube, M. et al. (2004) Hall scattering factors for electrons and holes in SiC, in Silicon Carbide—Recent Major Advances (eds W.J. Choyke, H. Matsunami and G. Pensl), Springer, p. 517.
  67. [67] Sze, S.M. and Ng, K.K. (2007) Physics of Semiconductor Devices, 3rd edn, John Wiley & Sons, Inc..
  68. [68] Bergman, J.P., Kordina, O. and Janzen, E. (1997) Time resolved spectroscopy of defects in SiC. Phys. Status Solidi A, 162, 65.
  69. [69] Klein, P.B. (2008) Carrier lifetime measurement in n- 4H-SiC epilayer. J. Appl. Phys., 103, 033702.
  70. [70] Kimoto, T., Hiyoshi, T., Hayashi, T. and Suda, J. (2010) Impacts of recombination at the surface and in the substrate on carrier lifetimes of n-type 4H-SiC epilayers. J. Appl. Phys., 108, 083721.
  71. [71] Galeckas, A., Linnros, J., Frischholz, M. and Grivickas, V. (2001) Optical characterization of excess carrier lifetime and surface recombination in 4H/6H–SiC. Appl. Phys. Lett., 79, 365.
  72. [72] Grivickas, P., Linnros, J. and Grivickas, V. (2001) Carrier diffusion characterization in epitaxial 4H–SiC. J. Mater. Res., 16, 524.
  73. [73] Hull, D. and Bacon, D.J. (2001) Introduction to Dislocations, 4th edn, Butterworth-Heinemann.
  74. [74] Katsuno, M., Ohtani, N., Takahashi, J. et al. (1999) Mechanism of molten KOH etching of SiC single crystals: comparative study with thermal oxidation. Jpn. J. Appl. Phys., 38, 4661.
  75. [75] Koga, K., Fujiwara, Y., Ueda, Y. and Yamaguchi, T. (1992) Growth and characterization of 6H-SiC bulk crystals by the sublimation method, in Amorphous and Crystalline Silicon Carbide IV (eds C.Y. Yang, M.M. Rahman and G.L. Harris), Springer-Verlag, p. 96.
  76. [76] Takahashi, J., Kanaya, M. and Fujiwara, Y. (1994) Sublimation growth of SiC single crystalline ingots on faces perpendicular to the (0001) basal plane. J. Cryst. Growth, 135, 61.
  77. [77] Yao, Y.Z., Ishikawa, Y., Sugawara, Y. et al. (2011) Molten KOH etching with Na2O2 additive for dislocation revelation in 4H-SiC epilayers and substrates. Jpn. J. Appl. Phys., 50, 075502.
  78. [78] Powell, J.A., Petit, J.B., Edgar, J.H. et al. (1991) Application of oxidation to the structural characterization of SiC epitaxial films. Appl. Phys. Lett., 59, 183.
  79. [79] Nakano, Y., Nakamura, T., Kamisawa, A. and Takasu, H. (2009) Investigation of pits formed at oxidation on 4H-SiC. Mater. Sci. Forum, 600–603, 377.
  80. [80] Suda, J., Shoji, H. and Kimoto, T. (2011) Origin of etch hillocks formed on on-aixs SiC (c05-math-0738 ) surfaces by molten KOH etching. Jpn. J. Appl. Phys., 50, 038002.
  81. [81] Yao, Y.Z., Ishikawa, Y., Sato, K. et al. (2012) Dislocation revelation from (c05-math-0739 ) carbon-face of 4H-SiC by using vaporized KOH at high temperature. Appl. Phys. Express, 5, 075601.
  82. [82] Bowen, D.K. and Tanner, B.K. (1998) High-Resolution X-Ray Diffractometry and Topography, Taylor & Francis Group.
  83. [83] Ohno, T., Yamaguchi, H., Kuroda, S. et al. (2004) Direct observation of dislocations propagated from 4H–SiC substrate to epitaxial layer by X-ray topography. J. Cryst. Growth, 260, 209.
  84. [84] Huang, X.R., Dudley, M., Vetter, W.M. et al. (1999) Direct evidence of micropipe-related pure superscrew dislocations in SiC. Appl. Phys. Lett., 74, 353.
  85. [85] Tsuchida, H., Kamata, I. and Nagano, M. (2007) Investigation of defect formation in 4H-SiC epitaxial growth by X-ray topography and defect selective etching. J. Cryst. Growth, 306, 254.
  86. [86] Tsuchida, H., Ito, M., Kamata, I. and Nagano, M. (2009) Fast epitaxial growth of 4H-SiC and analysis of defect transfer. Mater. Sci. Forum, 615–617, 67.
  87. [87] Jacobson, H., Birch, J., Yakimova, R. et al. (2002) Dislocation evolution in 4H-SiC epitaxial layers. J. Appl. Phys., 91, 6354.
  88. [88] Kamata, I., Nagano, M., Tsuchida, H. et al. (2009) Investigation of character and spatial distribution of threading edge dislocations in 4H-SiC epilayers by high-resolution topography. J. Cryst. Growth, 311, 1416.
  89. [89] Nakamura, D., Yamaguchi, S., Hirose, Y. et al. (2008) Direct determination of Burgers vector sense and magnitude of elementary dislocations by synchrotron white x-ray topography. J. Appl. Phys., 103, 013510.
  90. [90] Tanuma, R., Mori, D., Kamata, I. and Tsuchida, H. (2012) X-ray microbeam three-dimensional topography imaging and strain analysis of basal-plane dislocations and threading edge dislocations in 4H-SiC. Appl. Phys. Express, 5, 061301.
  91. [91] Tajima, M., Tanaka, M. and Hoshino, N. (2002) Characterization of SiC epitaxial wafers by photoluminescence under deep UV excitation. Mater. Sci. Forum, 389–393, 597.
  92. [92] Kamata, I., Tsuchida, H., Miyanagi, T. and Nakamura, T. (2005) Development of non-destructive in-house observation techniques for dislocations and stacking faults in SiC epilayers. Mater. Sci. Forum, 527–529, 415.
  93. [93] Camassel, J. and Juillaguet, S. (2008) Optical properties of as-grown and process-induced stacking faults in 4H-SiC. Phys. Status Solidi B, 245, 1337.
  94. [94] Feng, G., Suda, J. and Kimoto, T. (2008) Characterization of stacking faults in 4H-SiC epilayers by room-temperature microphotoluminescence mapping. Appl. Phys. Lett., 92, 221906.
  95. [95] Ito, M., Storasta, L. and Tsuchida, H. (2009) Development of a high rate 4H-SiC epitaxial growth technique achieving large-area uniformity. Mater. Sci. Forum, 600–603, 111.
  96. [96] Feng, G., Suda, J. and Kimoto, T. (2011) Nonradiative recombination at threading dislocations in 4H-SiC epilayers studied by micro-photoluminescence mapping. J. Appl. Phys., 110, 033525.
  97. [97] Isono, H., Tajima, M., Hoshino, N. and Sugimoto, H. (2009) Rapid characterization of SiC crystals by full-wafer photoluminescence imaging under below-gap excitation. Mater. Sci. Forum, 600–603, 545.
  98. [98] Stahlbush, R.E., Liu, K.X., Zhang, Q. and Sumakeris, J.J. (2007) Whole-wafer mapping of dislocations in 4H-SiC epitaxy. Mater. Sci. Forum, 556–557, 295.
  99. [99] Nagano, M., Kamata, I. and Tsuchida, T. (2013) Photoluminescence imaging and discrimination of threading dislocations in 4H-SiC epilayers. Mater. Sci. Forum, 740–742, 653.
  100. [100] Ichikawa, S., Kawahara, K., Suda, J. and Kimoto, T. (2012) Carrier recombination in n-type 4H-SiC epilayers with long carrier lifetimes. Appl. Phys. Express, 5, 101301.
  101. [101] Chung, S., Berechman, R.A., McCartney, M.R. and Skowronski, M. (2011) Electronic structure analysis of threading screw dislocations in 4H–SiC using electron holography. J. Appl. Phys., 109, 034906.
  102. [102] Liu, K.X., Zhang, X., Stahlbush, R.E. et al. (2009) Differences in emission spectra of dislocations in 4H-SiC epitaxial layers. Mater. Sci. Forum, 600–603, 345.
  103. [103] Caldwell, J.D., Giles, A., Lepage, D. et al. (2013) Experimental evidence for mobile luminescence center mobility on partial dislocations in 4H-SiC using hyperspectral electroluminescence imaging. Appl. Phys. Lett., 102, 242109.
  104. [104] Kitabatake, M. (2013) The International Conference on Silicon Carbide and Related Materials 2013, Miyazaki, Japan, We-1A-1. Electrical characteristics/reliability affected by defects analysed by the integrated evaluation platform for SiC epitaxial films
  105. [105] Milnes, A.G. (1973) Deep Impurities in Semiconductors, John Wiley & Sons, Inc.
  106. [106] Lang, D.V. (1974) Deep-level transient spectroscopy: A new method to characterize traps in semiconductors. J. Appl. Phys., 45, 3023.
  107. [107] Weiss, S. and Kassing, R. (1988) Deep Level Transient Fourier Spectroscopy (DLTFS) - A technique for the analysis of deep level properties. Solid State Electron., 31, 1733.
  108. [108] Dobaczewski, L., Kaczor, P., Hawkins, I.D. and Peaker, A.R. (1994) Laplace transform deep-level transient spectroscopic studies of defects in semiconductors. J. Appl. Phys., 76, 194.
  109. [109] Okushi, H. and Tokumaru, Y. (1981) Isothermal capacitance transient spectroscopy. Jpn. J. Appl. Phys., 20 (Suppl. 20-1), 261.
  110. [110] Lefevre, H. and Shultz, M. (1977) Double correlation technique (DDLTS) for the analysis of deep level profiles in semiconductors. Appl. Phys., 12, 45.
  111. [111] Reshanov, S.A., Pensl, G., Danno, K. et al. (2007) Effect of the Schottky barrier height on the detection of midgap levels in 4H-SiC by deep level transient spectroscopy. J. Appl. Phys., 102, 113702.
  112. [112] Weil, J.A. and Bolton, J.R. (2007) Electron Paramagnetic Resonance: Elementary Theory and Practical Applications, Wiley-Interscience.
  113. [113] Umeda, T., Isoya, J., Morishita, N. et al. (2004) EPR and theoretical studies of positively charged carbon vacancy in 4H-SiC. Phys. Rev. B, 70, 235212.
  114. [114] Itoh, H., Hayakawa, N., Nashiyama, I. and Sakuma, E. (1989) Electron spin resonance in electron-irradiated 3C-SiC. J. Appl. Phys., 66, 4529.
  115. [115] Son, N.T., Carlsson, P., Hassan, J. et al. (2006) Divacancy in 4H-SiC. Phys. Rev. Lett., 96, 055501.
  116. [116] Umeda, T., Isoya, J., Morishita, N. et al. (2009) Dicarbon antisite defect in n-type 4H-SiC. Phys. Rev. B, 79, 115211.
  117. [117] Isoya, J., Umeda, T., Mizuochi, N. et al. (2008) EPR identification of intrinsic defects in SiC. Phys. Status Solidi B, 245, 1298.
  118. [118] Umeda, T., Kosugi, R., Sakuma, Y. et al. (2012) MOS interface states: Similarity and dissimilarity from silicon. ECS Trans., 50, 305.
  119. [119] Lendenmann, H., Dahlquist, F., Johansson, N. et al. (2001) Long term operation of 4.5 kV PiN and 2.5 kV JBS diodes. Mater. Sci. Forum, 353–356, 727.
  120. [120] Bergman, J.P., Lendenmann, H., Nilsson, P.A. et al. (2001) Crystal defects as source of anomalous forward voltage increase of 4H-SiC diodes. Mater. Sci. Forum, 353–356, 299.
  121. [121] Lendenmann, H., Bergman, J.P., Dahlquist, F. and Hallin, C. (2003) Degradation in SiC bipolar devices: sources and consequences of electrically active dislocations in SiC. Mater. Sci. Forum, 433–436, 901.
  122. [122] Muzykov, P.G., Kennedy, R.M., Zhang, Q. et al. (2009) Physical phenomena affecting performance and reliability of 4H-SiC bipolar junction transistors. Microelectron. Reliab., 49, 32.
  123. [123] Skowronski, M. and Ha, S. (2006) Degradation of hexagonal silicon-carbide-based bipolar devices. J. Appl. Phys., 99, 011101.
  124. [124] Liu, J.Q., Skowronski, M., Hallin, C. et al. (2002) Structure of recombination-induced stacking faults in high-voltage SiC pn junctions. Appl. Phys. Lett., 80, 749.
  125. [125] Pirouz, P., Demenet, J.L. and Hong, M.H. (2001) On transition temperatures in the plasticity and fracture of semiconductors. Philos. Mag., 81, 1207.
  126. [126] Skowronski, M., Liu, J.Q., Vetter, W.M. et al. (2002) Recombination-enhanced defect motion in forward-biased 4H–SiC p-n diodes. J. Appl. Phys., 92, 4699.
  127. [127] Ha, S., Benamara, M., Skowronski, M. and Lendenmann, H. (2003) Core structure and properties of partial dislocations in silicon carbide pin diodes. Appl. Phys. Lett., 83, 4957.
  128. [128] Sridhara, S.G., Carlsson, F.H.C., Bergman, J.P. and Janzen, E. (2001) Luminescence from stacking faults in 4H SiC. Appl. Phys. Lett., 79, 3944.
  129. [129] Iwata, H., Lindefelt, U., Oberg, S. and Briddon, P.R. (2001) Localized electronic states around stacking faults in silicon carbide. Phys. Rev. B, 65, 033203.
  130. [130] Miyanagi, T., Tsuchida, H., Kamata, I. et al. (2006) Annealing effects on single Shockley faults in 4H-SiC. Appl. Phys. Lett., 89, 062104.
  131. [131] Caldwell, J.D., Stahlbush, R.E., Hobart, K.D. et al. (2007) Reversal of forward voltage drift in 4H-SiC pin diodes via low temperature annealing. Appl. Phys. Lett., 90, 143519.
  132. [132] Ha, S., Skowronski, M. and Lendenmann, H. (2004) Nucleation sites of recombination-enhanced stacking fault formation in silicon carbide p-i-n diodes. J. Appl. Phys., 96, 393.
  133. [133] Matare, H.F. (1971) Defect Electronics in Semiconductors, John Wiley & Sons, Inc.
  134. [134] Kimoto, T., Miyamoto, N. and Matsunami, H. (1999) Performance limiting surface defects in SiC epitaxial p-n junction diodes. IEEE Trans. Electron Devices, 46, 471.
  135. [135] Lindefelt, U. and Iwata, H. (2004) Electronic properties of stacking faults and thin cubic inclusions in SiC polytypes, in Silicon Carbide—Recent Major Advances (eds W.J. Choyke, H. Matsunami and G. Pensl), Springer, p. 89.
  136. [136] Fujiwara, H., Kimoto, T., Tojo, T. and Matsunami, H. (2005) Characterization of in-grown stacking faults in 4H–SiC (0001) epitaxial layers and its impacts on high-voltage Schottky barrier diodes. Appl. Phys. Lett., 87, 051912.
  137. [137] Neudeck, P.G. and Powell, J.A. (1994) Performance limiting micropipe defects in silicon carbide wafers. IEEE Electron Device Lett., 15, 63.
  138. [138] Neudeck, P.G., Huang, W. and Dudley, M. (1999) Study of bulk and elementary screw dislocation assisted reverse breakdown in low-voltage (< 250 V) 4H-SiC p+-n junction diodes. Part I: DC properties. IEEE Trans. Electron Devices, 46, 478.
  139. [139] Morisette, D.T. and Cooper, J.A. Jr. (2002) Impact of material defects on SiC Schottky barrier diodes. Mater. Sci. Forum, 389–393, 1133.
  140. [140] Saitoh, H., Kimoto, T. and Matsunami, H. (2004) Origin of leakage current in SiC Schottky barrier diodes at high temperature. Mater. Sci. Forum, 457–460, 997.
  141. [141] Wahab, Q., Ellison, A., Henry, A. et al. (2000) Influence of epitaxial growth and substrate-induced defects on the breakdown of 4H–SiC Schottky diodes. Appl. Phys. Lett., 76, 2725.
  142. [142] Tsuji, T., Izumi, S., Ueda, A. et al. (2002) Analysis of high leakage currents in 4H-SiC Schottky barrier diodes using optical beam-induced current measurements. Mater. Sci. Forum, 389–393, 1141.
  143. [143] Ohtani, N. (2011) Toward the reduction of performance-limiting defects in SiC epitaxial substrates. ECS Trans., 41, 253.
  144. [144] Fujiwara, H., Katsuno, T., Ishikawa, T. et al. (2012) Relationship between threading dislocation and leakage current in 4H-SiC diodes. Appl. Phys. Lett., 100, 242102.
  145. [145] Fujiwara, H., Naruoka, H., Konishi, M. et al. (2012) Impact of surface morphology above threading dislocations on leakage current in 4H-SiC diodes. Appl. Phys. Lett., 101, 042104.
  146. [146] Maranowski, M.M. and Cooper, J.A. Jr. (1999) Time-dependent-dielectric-breakdown measurements of thermal oxides on n-type 6H-SiC. IEEE Trans. Electron Devices, 46, 520.
  147. [147] Senzaki, J., Kojima, K. and Fukuda, K. (2004) Effects of n-type 4H-SiC epitaxial wafer quality on reliability of thermal oxides. Appl. Phys. Lett., 85, 6182.
  148. [148] Senzaki, J., Kojima, K., Kato, T. et al. (2006) Correlation between reliability of thermal oxides and dislocations in n-type 4H-SiC epitaxial wafers. Appl. Phys. Lett., 89, 022909.
  149. [149] Tanimoto, S. (2006) Impact of dislocations on gate oxide in SiC MOS devices and high reliability ONO dielectrics. Mater. Sci. Forum, 527–529, 955.
  150. [150] Senzaki, J., Shimozato, A., Kojima, K. et al. (2012) Challenges of high-performance and high-reliablity in SiC MOS structures. Mater. Sci. Forum, 717–720, 703.
  151. [151] Sze, S.M. (2002) Semiconductor Devices, Physics and Technology, 2nd edn, John Wiley & Sons, Inc.
  152. [152] Dalibor, T., Pensl, G., Matsunami, H. et al. (1997) Deep defect centers in silicon carbide monitored with deep level transient spectroscopy. Phys. Status Solidi A, 162, 199.
  153. [153] Hemmingsson, C., Son, N.T., Kordina, O. et al. (1997) Deep level defects in electron-irradiated 4H SiC epitaxial layers. J. Appl. Phys., 81, 6155.
  154. [154] Danno, K. and Kimoto, T. (2007) Deep level transient spectroscopy on as-grown and electron-irradiated p-type 4H-SiC epilayers. J. Appl. Phys., 101, 103704.
  155. [155] Alfieri, G. and Kimoto, T. (2013) Resolving the EH6/7 level in 4H-SiC by Laplace-transform deep level transient spectroscopy. Appl. Phys. Lett., 102, 152108.
  156. [156] Hemmingsson, C.G., Son, N.T., Ellison, A. et al. (1998) Negative-U centers in 4H silicon carbide. Phys. Rev. B, 58, R10119.
  157. [157] Grundmann, M. (2006) The Physics of Semiconductors Chapter 7, Springer.
  158. [158] Danno, K. and Kimoto, T. (2006) Investigation of deep levels in n-type 4H-SiC epilayers irradiated with low-energy electrons. J. Appl. Phys., 100, 113728.
  159. [159] Sasaki, S., Kawahara, K., Feng, G. et al. (2011) Major deep levels with the same microstructures observed in n-type 4H-SiC and 6H-SiC. J. Appl. Phys., 109, 013705.
  160. [160] Hornos, T., Gali, A. and Svensson, B.G. (2011) Negative-U system of carbon vacancy in 4H-SiC. Mater. Sci. Forum, 679–680, 261.
  161. [161] Son, N.T., Trinh, X.T., Løvlie, L.S. et al. (2012) Large-scale electronic structure calculations of vacancies in 4H-SiC using the Heyd-Scuseria-Ernzerhof screened hybrid density functional. Phys. Rev. Lett., 109, 187603.
  162. [162] Kawahara, K., Trinh, X.T., Son, N.T. et al. (2013) Investigation on origin of Z1/2 center in SiC by deep level transient spectroscopy and electron paramagnetic resonance. Appl. Phys. Lett., 102, 112106.
  163. [163] Kawahara, K., Trinh, X.T., Son, N.T. et al. (2014) Quantitative comparison between Z1/2 center and carbon vacancy in 4H-SiC. J. Appl. Phys, 115, 143705.
  164. [164] Alfieri, G. and Kimoto, T. (2009) Capacitance spectroscopy study of midgap levels in n-type SiC polytypes. Mater. Sci. Forum, 615–617, 389.
  165. [165] Storasta, L., Bergman, J.P., Janzen, E. et al. (2004) Deep levels created by low energy electron irradiation in 4H-SiC. J. Appl. Phys., 96, 4909.
  166. [166] Kaneko, H. and Kimoto, T. (2011) Formation of a semi-insulating layer in n-type 4H-SiC by electron irradiation. Appl. Phys. Lett., 98, 262106.
  167. [167] Zhang, J., Storasta, L., Bergman, J.P. et al. (2003) Electrically active defects in n-type 4H-silicon carbide grown in a vertical hot-wall reactor. J. Appl. Phys., 93, 4708.
  168. [168] Kimoto, T., Hashimoto, K. and Matsunami, H. (2003) Effects of C/Si ratio in chemical vapor deposition of 4H-SiC(c05-math-0740 ) and (c05-math-0741 ). Jpn. J. Appl. Phys., 42, 7294.
  169. [169] Alfieri, G., Monakhov, E.V., Svensson, B.G. and Linnarsson, M.K. (2005) Annealing behavior between room temperature and 2000 °C of deep level defects in electron-irradiated n-type 4H silicon carbide. J. Appl. Phys., 98, 043518.
  170. [170] Zippelius, B., Suda, J. and Kimoto, T. (2012) High temperature annealing of n-type 4H-SiC: Impact on intrinsic defects and carrier lifetime. J. Appl. Phys., 111, 033515.
  171. [171] Klein, P.B., Shanabrook, B.V., Huh, S.W. et al. (2006) Lifetime-limiting defects in 4H-SiC epilayers. Appl. Phys. Lett., 88, 052110.
  172. [172] Danno, K., Nakamura, D. and Kimoto, T. (2007) Investigation of carrier lifetime in 4H-SiC epilayers and lifetime control by electron irradiation. Appl. Phys. Lett., 90, 202109.
  173. [173] Storasta, L. and Tsuchida, H. (2007) Reduction of traps and improvement of carrier lifetime in 4H-SiC epilayers by ion implantation. Appl. Phys. Lett., 90, 062116.
  174. [174] Storasta, L., Tsuchida, H., Miyazawa, T. and Ohshima, T. (2008) Enhanced annealing of the Z1/2 defect in 4H–SiC epilayers. J. Appl. Phys., 103, 013705.
  175. [175] Dunham, S.T. and Plummer, J.D. (1986) Point-defect generation during oxidation of silicon in dry oxygen. I. Theory. J. Appl. Phys., 59, 2541.
  176. [176] Hiyoshi, T. and Kimoto, T. (2009) Reduction of deep levels and improvement of carrier lifetime in n-type 4H-SiC by thermal oxidation. Appl. Phys. Express, 2, 041101.
  177. [177] Hiyoshi, T. and Kimoto, T. (2009) Elimination of the major deep levels in n- and p-type 4H-SiC by two-step thermal treatment. Appl. Phys. Express, 2, 091101.
  178. [178] Kawahara, K., Suda, J. and Kimoto, T. (2012) Analytical model for reduction of deep levels in SiC by thermal oxidation. J. Appl. Phys., 111, 053710.
  179. [179] Bockstedte, M., Mattausch, A. and Pankratov, O. (2004) Ab initio study of the annealing of vacancies and interstitials in cubic SiC: Vacancy-interstitial recombination and aggregation of carbon interstitials. Phys. Rev. B, 69, 235202.
  180. [180] Hong, J.D., Davis, R.F. and Newbury, D.E. (1981) Self-diffusion of silicon-30 in α-SiC single crystals. J. Mater. Sci., 16, 2485.
  181. [181] Kawahara, K., Suda, J. and Kimoto, T. (2013) Deep levels generated by thermal oxidation in p-type 4H-SiC. J. Appl. Phys., 113, 033705.
  182. [182] Kawahara, K., Suda, J. and Kimoto, T. (2013) Deep levels generated by thermal oxidation in n-type 4H-SiC. Appl. Phys. Express, 6, 051301.
  183. [183] Danno, K., Hori, T. and Kimoto, T. (2007) Impacts of growth parameters on deep levels in n-type 4H-SiC. J. Appl. Phys., 101, 053709.
  184. [184] Lebedev, A.A. (1999) Deep level centers in silicon carbide: A review. Semiconductors, 33, 107.
  185. [185] Zywietz, A., Furthmüller, J. and Bechstedt, F. (1999) Vacancies in SiC: Influence of Jahn-Teller distortions, spin effects, and crystal structure. Phys. Rev. B, 59, 15166.
  186. [186] Torpo, L., Marlo, M., Staab, T.E.M. and Nieminen, R.M. (2001) Comprehensive ab initio study of properties of monovacancies and antisites in 4H-SiC. J. Phys. Condens. Matter, 13, 6203.
  187. [187] Troffer, T., Schadt, M., Frank, T. et al. (1997) Doping of SiC by implantation of boron and aluminum. Phys. Status Solidi A, 162, 277.
  188. [188] Duijn-Arnold, A., Ikoma, T., Poluektov, O.G. et al. (1998) Doping of SiC by implantation of boron and aluminum. Phys. Rev. B, 57, 1607.
  189. [189] Negoro, Y., Kimoto, T. and Matsunami, H. (2005) Carrier compensation near tail region in aluminum- or boron-implanted 4H-SiC(0001). J. Appl. Phys., 98, 043709.
  190. [190] Dalibor, T., Trageser, H., Pensl, G. et al. (1999) Oxygen in silicon carbide: Shallow donors and deep acceptors. Mater. Sci. Eng., B, 61–62, 454.
  191. [191] Klettke, O., Pensl, G., Kimoto, T. and Matsunami, H. (2001) Oxygen-related defect centers observed in 4H/6H-SiC epitaxial layers grown under CO2 ambient. Mater. Sci. Forum, 353–356, 459.
  192. [192] Langer, J.M. and Heinrich, H. (1985) Deep-level impurities: a possible guide to prediction of band-edge discontinuities in semiconductor heterojunctions. Phys. Rev. Lett., 55, 1414.
  193. [193] Kunzer, M., Kaufmann, U., Maier, K. and Schneider, J. (1995) Magnetic circular dichroism and electron spin resonance of the A acceptor state of vanadium, V3+, in 6H-SiC. Mater. Sci. Eng., B, 29, 118.
  194. [194] Achtziger, N. and Witthuhn, W. (2004) Radiotracer deep level transient spectroscopy, in Silicon Carbide—Recent Major Advances (eds W.J. Choyke, H. Matsunami and G. Pensl), Springer, p. 537.
  195. [195] Beyer, F., Hemmingsson, C., Leone, S. et al. (2011) Deep levels in iron doped n- and p-type 4H-SiC. J. Appl. Phys., 110, 123701.
  196. [196] Danno, K., Saitoh, H., Seki, A. et al. (2012) Diffusion of transition metals in 4H-SiC and trials of impurity gettering. Appl. Phys. Express, 5, 031301.
  197. [197] Tawara, T., Tsuchida, H. and Izumi, S. (2004) Evaluation of free carrier lifetime and deep levels of the thick 4H-SiC epilayers. Mater. Sci. Forum, 457–460, 565.
  198. [198] Kimoto, T., Danno, K. and Suda, J. (2008) Lifetime-killing defects in 4H-SiC epilayers and lifetime control by low-energy electron irradiation. Phys. Status Solidi B, 245, 1327.
  199. [199] Reshanov, S.A., Bartsch, W., Zippelius, B. and Pensl, G. (2009) Lifetime investigations of 4H-SiC pin power diodes. Mater. Sci. Forum, 615–617, 699.
  200. [200] Noborio, M., Suda, J., Beljakowa, S. et al. (2009) 4H-SiC MISFETs with nitrogen-containing insulators. Phys. Status Solidi A, 206, 2374.
  201. [201] Miyazawa, T., Ito, M. and Tsuchida, H. (2010) Evaluation of long carrier lifetimes in thick 4H silicon carbide epitaxial layers. Appl. Phys. Lett., 97, 202106.
  202. [202] Hayashi, T., Asano, T., Suda, J. and Kimoto, T. (2011) Temperature and injection level dependencies and impact of thermal oxidation on carrier lifetimes in p-type and n-type 4H-SiC epilayers. J. Appl. Phys., 109, 114502.
  203. [203] Hayashi, T., Asano, K., Suda, J. and Kimoto, T. (2012) Enhancement and control of carrier lifetimes in p-type 4H-SiC epilayers. J. Appl. Phys., 112, 064503.
  204. [204] Maximenko, S.I., Freitas, J.A. Jr., Klein, P.B. et al. (2009) Cathodoluminescence study of the properties of stacking faults in 4H-SiC homoepitaxial layers. Appl. Phys. Lett., 94, 092101.
  205. [205] Hassan, J. and Bergman, J.P. (2009) Influence of structural defects on carrier lifetime in 4H-SiC epitaxial layers: Optical lifetime mapping. J. Appl. Phys., 105, 123518.
..................Content has been hidden....................

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