Chapter 10
Optimization and Comparison of Power Devices

10.1 Blocking Voltage and Edge Terminations for SiC Power Devices

An essential requirement of all semiconductor power devices is the ability to withstand a large terminal voltage with minimal leakage current in the off state. The maximum terminal voltage a device can withstand is called the blocking voltage. The blocking voltage is generally determined by material properties as well as device design. The limiting mechanism may be any of the following: (i) punch-through of the base region in a MOSFET (metal-oxide-semiconductor field effect transistor), BJT (bipolar junction transistor), IGBT (insulated-gate bipolar transistor), or thyristor; (ii) avalanche breakdown in a reverse-biased pn or Schottky junction, either as a discrete rectifier or as a part of a switching transistor or thyristor; (iii) excessive leakage current in a reverse-biased pn or Schottky junction; or (iv) excessive electric field in the oxide of an MOS-based power device such as a MOSFET or IGBT.

Punch-through can be avoided by making the doping-thickness product of the base region large enough that the base cannot be completely depleted before the onset of avalanche breakdown. This requires

10.1 equation

where c10-math-0002 and c10-math-0003 represent the doping and width of the base region and c10-math-0004 is the critical field for avalanche breakdown, to be discussed below. Since this restriction is typically observed for all power switching devices, punch-through will not be considered further.

In an MOS-based device such as a power MOSFET or IGBT, the blocking voltage is often limited by the oxide field. As discussed in Sections 8.2.6 and 8.2.11, the oxide field must be kept below about c10-math-0005 for long-term device reliability. Care must be exercised in the design of MOSFETs and IGBTs to make sure the oxide field does not exceed this value prior to avalanche breakdown of the blocking junction. This typically requires numerical simulations due to the two-dimensional nature of the fields, particularly in trench-gate or UMOS structures.

In well-designed devices, the blocking voltage is ultimately limited by avalanche breakdown of the reverse-biased blocking junction. The blocking voltage is reduced by two-dimensional field crowding at the edges of the device, which can be mitigated by the use of special edge terminations. In the sections below, we will first consider avalanche breakdown in planar SiC junctions using a one-dimensional analysis. We will then discuss two-dimensional field crowding and show how it can be minimized by different edge termination techniques.

10.1.1 Impact Ionization and Avalanche Breakdown

Avalanche breakdown is caused by the impact ionization of electrons and holes in a high-field region. The impact ionization process can be understood by considering the behavior of electrons and holes as they move through the crystal. As charged entities, electrons and holes are accelerated by the electric field, and their kinetic energy increases until they undergo a collision. The collision is a scattering event that takes the electron (or hole) into a lower energy state, with its energy typically transferred to the crystal lattice as heat. Immediately following the scattering event, the electron or hole is again accelerated, and the process is repeated as holes and electrons transit the high-field region. If the field is sufficiently high, the electron or hole may acquire enough kinetic energy between collisions that the energy liberated is sufficient to break a covalent bond, creating a new hole–electron pair. This process is called impact ionization.

The number of impact ionization events initiated by an electron (or hole) per unit length traveled is known as the ionization rate for electrons c10-math-0006 or the ionization rate for holes c10-math-0007. c10-math-0008 and c10-math-0009 are strong functions of the electric field. Figure 10.1 shows the ionization rates for transport along the c-axis in 4H-SiC at room temperature, measured by Konstantinov et al. [1] This data can be represented by the empirical expressions [2]

10.2 equation

and

10.3 equation
c10f001

Figure 10.1 Ionization rate coefficients in 4H-SiC measured at room temperature ( [1] reproduced with permission from AIP Publishing LLC).

Figure 10.2 compares the ionization rates for silicon and 4H-SiC. The ionization rates for 4H-SiC at a given field are orders-of-magnitude lower than for silicon.

c10f002

Figure 10.2 Comparison of ionization rates in silicon and 4H-SiC at room temperature. At a given field, the rates for SiC are orders-of-magnitude below those of silicon.

Avalanche breakdown in 4H-SiC is typically initiated by holes, since the ionization rates are higher for holes than for electrons. Figure 10.3 shows the ionization rates for holes in 4H-SiC for several temperatures [3]. The lines represent the empirical equation

where c10-math-0013 is the absolute temperature. The decrease in ionization rates with temperature occurs because of increased phonon scattering, making it less likely that a hole can acquire enough kinetic energy between collisions to create a hole–electron pair.

c10f003

Figure 10.3 Temperature dependence of the hole ionization rate in 4H-SiC. The lines are generated using empirical Equation 10.4 ([3] reproduced with permission from Trans Tech Publications).

We will now apply the ionization rate expressions to calculate the breakdown voltage of a reverse-biased c10-math-0014 one-sided step junction, such as shown in Figure 7.1. The depletion region extends from c10-math-0015 to c10-math-0016 and the electric field decreases from a peak value at c10-math-0017 to zero at c10-math-0018. Any electrons and holes in the depletion region are subject to an electric field and can potentially initiate impact ionization events. Consider the cross-section of the depletion region shown in Figure 10.4. We assume hole current c10-math-0019 is entering at c10-math-0020 and hole current c10-math-0021 is flowing at point c10-math-0022. Impact ionization in the differential slice at c10-math-0023 creates additional holes and electrons, and the increase in hole current across the c10-math-0024 region is

10.5 equation

which we can write as

c10f004

Figure 10.4 One-dimensional cross-section of a high-field region, illustrating generation of new electron–hole pairs by impact ionization.

Since c10-math-0027 is not a function of c10-math-0028, we can write Equation 10.6 as

Since c10-math-0030 and c10-math-0031 are functions of electric field and the electric field is a function of position, Equation 10.7 can be written in the form

where c10-math-0033, and c10-math-0034 are functions of c10-math-0035. The differential Equation 10.8 has the formal solution

Substituting for c10-math-0037, and c10-math-0038 in Equation 10.9 yields

We define the hole multiplication factor c10-math-0040 as the fractional increase in hole current across the depletion region due to impact ionization, c10-math-0041. Since we assumed no electron current at c10-math-0042 and we can set c10-math-0043. Under these assumptions, if we now evaluate Equation 10.10 at c10-math-0044 we can write

10.11 equation

The complicated fraction on the right-hand side must be equal to 1, so c10-math-0046 is given by

10.12 equation

The exponential factor in the second term in the denominator can be rewritten using the identity

10.13 equation

resulting in

10.14 equation

Avalanche breakdown occurs when the multiplication factor c10-math-0050 goes to infinity, or when

Equation 10.15 is known as the ionization integral for holes. The above analysis could also have been conducted assuming only an electron current entering at c10-math-0052. In this case we would have defined an electron multiplication factor c10-math-0053, and avalanche breakdown would occur when c10-math-0054 goes to infinity, or when

Considering that both holes and electrons are present in the device, avalanche breakdown will occur at the lowest voltage where either Equation 10.15 or 10.16 is satisfied.

It is customary to designate the peak field c10-math-0056 at the onset of breakdown as the critical field for avalanche breakdown, c10-math-0057. c10-math-0058 can be computed from the ionization integrals by an iterative procedure, as follows: For a given doping of the c10-math-0059 layer, select a value of reverse voltage and calculate the ionization integrals using Equations 10.15 and 10.16. If both integrals are less than 1, chose a slightly higher voltage and repeat the calculation. The lowest voltage for which either Equation 10.15 or 10.16 is satisfied is the breakdown voltage for that doping, and the peak field at the junction is the critical field for that doping. The critical field for a one-sided step junction in 4H-SiC at room temperature can be approximated using an empirical expression given by Konstantinov et al. [1],

where c10-math-0061 is the doping of the lightly-doped side of the junction. Figure 10.5 shows the critical field in silicon and 4H-SiC as a function of doping. The critical field is only a weak function of doping, and at doping levels below c10-math-0062 the critical field in 4H-SiC is approximately seven times higher than in silicon. Since the power-device figure of merit scales inversely as the cube of critical field (see Section 7.1), this represents about a c10-math-0063 for 4H-SiC.

c10f005

Figure 10.5 Critical field for room temperature avalanche breakdown in a 4H-SiC one-sided step junction, as given by Equation 10.17. Also shown for comparison is the critical field for silicon.

As might be expected from the temperature dependence of the ionization rates, the avalanche breakdown voltage in 4H-SiC increases with temperature, as shown in Figure 10.6 [4]. A positive temperature coefficient of breakdown is desirable, since devices with negative temperature coefficient of breakdown are potentially unstable.

c10f006

Figure 10.6 Temperature dependence of avalanche breakdown voltage in 4H and 6H-SiC. The temperature coefficient of breakdown is positive in 4H-SiC and negative in 6H-SiC ([4] reproduced with permission from AIP Publishing LLC).

It is important to recognize that the critical field in Equation 10.17 was calculated assuming a non-punch-through structure such as in Figure 7.1. In the punch-through structure of Figure 7.2, evaluation of the ionization integral yields a critical field that depends not only on doping, but also on the thickness of the lightly-doped region. Figure 10.7 shows the critical field in 4H-SiC calculated using the ionization integrals for both punch-through and non-punch-through structures [2]. The dashed line is the non-punch-through critical field given by Equation 10.17. In punch-through designs, the critical field for breakdown is higher than the value given by Equation 10.17.

c10f007

Figure 10.7 Critical field for room temperature avalanche breakdown in a 4H-SiC one-sided step junction with a punch-through structure. The critical field for a non-punch-through structure, given by Equation 10.17, is shown as a dashed line ([2] reproduced with permission from Dallas T. Morisette).

10.1.2 Two-Dimensional Field Crowding and Junction Curvature

In most of our discussions to this point, we have considered only one-dimensional slices within our devices. This allows us to calculate carrier densities, current flows, and electrostatic potentials using one-dimensional analyses. In some cases this approach gives quantitatively accurate answers, but it often happens that the two-dimensional (or three-dimensional) nature of real devices necessitates a computer simulation. This is certainly true in the calculation of blocking voltage. The results in the preceding subsection were obtained using a one-dimensional analysis, but in real devices the blocking voltage is invariably limited by two-dimensional field crowding at the periphery of the device. Next, we will discuss how to evaluate the field crowding and consider techniques to mitigate it.

Consider the cylindrical c10-math-0064 one-sided step junction illustrated in Figure 10.8. The c10-math-0065 region has radius c10-math-0066 and the depletion edge has radius c10-math-0067. Poisson's equation in cylindrical coordinates is

10.18 equation

where c10-math-0069 is the radial coordinate, c10-math-0070, and c10-math-0071 is the charge per unit volume in the depletion region. Integration with respect to c10-math-0072 yields

10.19 equation
c10f008

Figure 10.8 Cross-section of a c10-math-0074 one-sided step junction with cylindrical geometry, such as occurs at the edges of real structures. c10-math-0075 is the radius of the junction and c10-math-0076 is the radius of the depletion region edge.

Noting that c10-math-0077 and choosing the constant of integration such that c10-math-0078, we can write

The peak field occurs at the metallurgical junction at c10-math-0080,

Integrating Equation 10.20 with respect to c10-math-0082 to calculate the potential c10-math-0083, and setting c10-math-0084, we obtain

10.22 equation

The total voltage across the depletion region is the potential difference c10-math-0086,

The peak field can be plotted as a function of reverse voltage by choosing successively larger values of c10-math-0088 and evaluating Equations 10.21 and 10.23. Figure 10.9 shows the effect of junction curvature on the peak field for several values of c10-math-0089 at an assumed doping of c10-math-0090. The peak field of a planar junction of the same doping is also shown for comparison. Reducing the junction radius c10-math-0091 increases the peak field at a given reverse voltage, and this will significantly reduce the breakdown voltage. Because of this effect, the blocking voltage of practical devices is normally limited by edge breakdown unless special edge termination techniques are employed.

c10f009

Figure 10.9 Peak field at the cylindrical junction as a function of reverse voltage, with junction radius as a parameter.

Edge terminations used in SiC devices fall into five general classes: trench isolation, beveled junctions, junction termination extension (JTE), floating field rings (FFRs), and multiple floating zone (MFZ-) or space-modulated (SM-) JTE. We will discuss each of these below.

10.1.3 Trench Edge Terminations

Trench isolation is useful when the reverse voltage is below about 2 kV and the depletion width at breakdown is on the order of c10-math-0092 or less. Under these conditions it is possible to etch an isolation trench extending c10-math-0093 through the lightly-doped drift region, as illustrated in Figure 10.10. This simple technique is often sufficient because it tends to linearize field lines near the c10-math-0094 junction where the field is highest. Although field crowding occurs at the corner of the trench, the field here is considerably lower than the peak field, reducing the negative impact of field crowding. A high-quality passivation layer is required to control surface charges and prevent surface leakage.

c10f010

Figure 10.10 Cross-section of a c10-math-0095 one-sided step junction terminated by trench isolation.

10.1.4 Beveled Edge Terminations

At higher blocking voltages, an effective way to provide edge termination is the use of a beveled junction profile. A positive bevel is one in which the junction area decreases with distance into the lightly-doped side, as illustrated in Figure 10.11 for a c10-math-0096 junction. The requirement for charge balance on either side of the junction causes an increase in the depletion width on the lightly-doped side near the beveled surface. In this case, the charge balance is restored by replacing the positive charge missing from area c10-math-0097 by a roughly equal additional charge c10-math-0098. As a result, the depletion width along the beveled surface c10-math-0099 is larger than the depletion width c10-math-0100 in the bulk. Since the potential drop across the junction is the same everywhere, the electric field along the surface is reduced relative to that in the bulk. Numerical simulations confirm that the surface field is reduced below the bulk field for all bevel angles c10-math-0101 [5]. In practice, most SiC power devices are fabricated with the heavily-doped layer on top, making it difficult to achieve a positive bevel in real devices.

c10f011

Figure 10.11 Charge distribution in a c10-math-0102 junction with a positive bevel.

A negative bevel is one in which the junction area increases with distance into the lightly-doped side, as illustrated in Figure 10.12 for an c10-math-0103 junction. In this case the depletion region on the lightly-doped side shrinks near the surface to balance the missing charge in the heavily-doped side. This tends to make the depletion width along the surface c10-math-0104 smaller than the depletion width c10-math-0105 in the bulk, increasing the surface field above the bulk field. However, at very small bevel angles the shallow slope makes the increase in depletion width on the heavily-doped side exceed the decrease in depletion width on the lightly-doped side, as illustrated in the lower part of Figure 10.12. When this occurs, the peak surface field is once again lower than the peak field in the bulk.

c10f012

Figure 10.12 Charge distribution in an c10-math-0106 junction with a negative bevel.

The exact conditions under which a negative bevel reduces the surface field can be predicted by a two-dimensional numerical solution of Poisson's equation. Adler and Temple have studied negative bevels in silicon junctions where the heavily-doped layer is formed by diffusion [6]. After extensive simulations, they found that for negative beveled junctions the normalized breakdown voltage (as a fraction of the ideal breakdown voltage) can be represented by a universal curve that depends on an effective bevel angle c10-math-0107 given by

10.24 equation

Here c10-math-0109 is the actual bevel angle, c10-math-0110 is the depletion width on the lightly-doped side at breakdown, and c10-math-0111 is the depletion width on the heavily-doped side at breakdown. Figure 10.13 shows normalized breakdown voltage as a function of effective bevel angle. For sufficiently small angles the breakdown approaches the ideal planar-junction value. However, these simulations were carried out for graded, diffused junctions with a complementary error function doping profile, and the results may not be accurate for the abrupt junctions typically formed in SiC.

c10f013

Figure 10.13 Normalized breakdown voltage as a function of effective bevel angle c10-math-0112 for a negatively-beveled silicon junction with a graded doping profile ( [6] reproduced with permission from IEEE).

Figure 10.14 shows the maximum field at the surface and the maximum field in the bulk material, normalized to the ideal breakdown field, as a function of effective bevel angle [6]. At small bevel angles the maximum surface field is well below the ideal (planar) breakdown field, but the maximum field in the bulk is slightly higher than the ideal breakdown field. This means that avalanche breakdown first occurs in the bulk near the surface, rather than at the surface itself, and the breakdown voltage is therefore lower than the ideal (planar) value. As the effective bevel angle is increased, both the maximum surface field and the maximum bulk field increase and the breakdown voltage is further reduced.

c10f014

Figure 10.14 Maximum field at the surface and maximum field in the bulk material as a function of effective bevel angle for negatively-beveled silicon junctions with a graded doping profile ( [6] reproduced with permission from IEEE).

10.1.5 Junction Termination Extensions (JTEs)

The third form of edge termination used in SiC is junction termination extension (JTE) [7]. This termination consists of one or more concentric p-type rings of carefully-controlled dopant concentration surrounding the main junction, as shown in Figure 10.15. The key requirement is that the total dopant concentration per unit area in each ring be low enough that the ring will be completely depleted before avalanche breakdown occurs at the edge of the ring. The exposed acceptor atoms in the depleted rings then terminate field lines that otherwise would crowd into the corner of the main junction. The field lines in this situation are illustrated schematically in Figure 10.16.

c10f015

Figure 10.15 Cross-section of a c10-math-0113 one-sided step junction protected by a two-zone junction termination extension (JTE).

c10f016

Figure 10.16 Illustration of field lines in a two-zone JTE termination under reverse bias.

Because of the two-dimensional nature of the fields, optimization of JTE ring design is best performed by computer simulations. Figure 10.17 shows breakdown voltage as a function of dose for single-zone JTE terminations of various widths on a c10-math-0114 SiC drift region doped c10-math-0115 [8]. The breakdown voltage peaks at the dose where the ring becomes depleted just before reaching breakdown. For higher doses, the ring does not fully deplete and breakdown occurs at the outer edge of the ring. For lower doses, the ring depletes without breaking down, and breakdown occurs at the edge of the main junction.

c10f017

Figure 10.17 Blocking voltage as a function of ring dose (dopant concentration per unit area) for single-zone JTE terminations with ring width as a parameter. The c10-math-0116 drift region has a theoretical planar breakdown voltage of 13.8 kV ([8] reproduced with permission from Imran A. Khan).

For a given ring width, the breakdown voltage falls rapidly for doses higher than the optimum dose. For this reason, it is customary to select a dose that is approximately 75% of the optimum, to allow for variation in implant activation percentage during processing. The maximum breakdown voltage increases with ring width until the width is about twice the thickness of the depletion region at breakdown.

The narrow dose window to achieve a high breakdown voltage can be mitigated by the use of multiple JTE rings. Figure 10.18 shows the breakdown voltage for a three-zone JTE system on a c10-math-0117 drift region doped c10-math-0118 [9]. The dose of the inner ring is fixed at three times that of the outer ring, and the dose of the middle ring is twice that of the outer ring. The breakdown voltage exhibits three peaks as a function of dose. These correspond to transitions in the location of the breakdown point within the structure. At high doses, none of the rings deplete and breakdown occurs at the edge of the outer ring. As the outer ring dose falls below c10-math-0119, this ring becomes completely depleted before breaking down and the breakdown point shifts to the edge of the middle ring. When the outer ring dose falls below c10-math-0120, corresponding to a middle ring dose of c10-math-0121, the middle ring also depletes before breaking down and breakdown shifts to the edge of the inner ring. When the outer ring dose falls below c10-math-0122, corresponding to an inner ring dose of c10-math-0123, the inner ring depletes and breakdown moves to the edge of the main junction. By including multiple rings, the range of dose over which a high breakdown voltage is achieved is widened considerably.

c10f018

Figure 10.18 Blocking voltage as a function of ring dose for a three-zone JTE termination. The c10-math-0124 drift region is c10-math-0125 thick, doped c10-math-0126, with a theoretical planar breakdown voltage of 12 kV ([9] reproduced with permission from Trans Tech Publications).

Simulations indicate that the depth and dopant profile of the JTE implant have little effect on performance [9], but surface charge can play an important role and should be controlled by the use of a high-quality oxide passivation layer.

10.1.6 Floating Field-Ring (FFR) Terminations

Floating field-ring (FFR) terminations consist of a series of isolated concentric c10-math-0128 rings surrounding the main junction, and are usually formed in the same processing step as the main junction so that no additional processing is required. Since the rings are heavily doped, the performance of a floating field-ring system does not require precise control of activated implant dose.

Figure 10.19 illustrates a four-ring FFR termination along with equipotential lines under reverse bias. As the reverse voltage is increased from zero, the depletion region of the main junction gradually expands until it reaches the first field ring. This ring then acts as an equipotential region and the depletion region continues to expand outward from this ring. The equipotential lines in the figure may be taken as representative of the successive positions of the depletion edge as the reverse voltage is gradually increased. The effect of the FFRs is to spread out the potential distribution laterally along the surface, reducing the lateral electric field that would otherwise initiate avalanche breakdown at the main junction.

c10f019

Figure 10.19 Cross-section of a c10-math-0127 one-sided step junction protected by four floating field rings (FFRs). Several equipotential lines under reverse bias are illustrated.

Figure 10.20 shows electric field and potential along the surface under 1975 V reverse bias for a four-ring FFR termination on a c10-math-0129 drift region doped c10-math-0130 [10]. The highest field values occur at the outer edges of the field rings. The ring system spreads the potential laterally along the surface, as seen from the potential plot.

c10f020

Figure 10.20 Electric field (a) and electrostatic potential (b) at the surface of a four-ring FFR termination on a c10-math-0131 drift region doped c10-math-0132. The reverse voltage is 1.975 kV and the ring parameters are c10-math-0133, and c10-math-0134 ( [10], reproduced with permission from James A. Cooper).

Since the field rings are all heavily doped, doping is not a design variable, but the number of rings, the width of each ring, and the spacing of each ring from the next inner ring are all design parameters. To be effective, the ring system should extend laterally at least twice the depth of the depletion region at breakdown. For high-voltage devices it is not unusual to incorporate several tens of concentric rings, and this introduces an inordinate number of free design parameters. Therefore, it is helpful to adopt a systematic method of specifying ring widths and spacings. For example, one approach is to require that the width and spacing of every ring have the same ratio c10-math-0135, and that the width (and spacing) increase with successive rings by a fixed expansion ratio defined as

10.25 equation

In this implementation, the period c10-math-0137 also increases with successive rings by the expansion ratio c10-math-0138. Under these constraints, one can evaluate FFR system performance as a function of the initial spacing of the first ring c10-math-0139, the c10-math-0140 ratio, the expansion ratio c10-math-0141, and the total number of rings in the system. Figure 10.21 shows breakdown voltage obtained by numerical simulation for a number of different ring systems on a c10-math-0142 SiC n-type drift layer doped c10-math-0143 [10]. This layer has a theoretical planar breakdown voltage of 3.5 kV. In the plot, each point denotes a specific number of rings and total width of the ring system for a particular design, and the expressions in parentheses give the initial spacing c10-math-0144, expansion ratio c10-math-0145, and width-to-spacing ratio c10-math-0146.

c10f021

Figure 10.21 Blocking voltage versus total ring width for a variety of FFR terminations on a c10-math-0147 drift region doped c10-math-0148 [9]. Each point represents a possible ring system, with the total width of the ring system given on the horizontal axis ( [10], reproduced with permission from James A. Cooper).

Although no specific algorithm for optimum design emerges, we can make some general observations. First, increasing the total width of the ring system increases the breakdown voltage, but the improvement saturates when the ring system extends beyond the lateral depletion width at breakdown. Second, comparing curve (a) to (c) and curve (b) to (d) we conclude that an expansion ratio of 5% (1.05) gives a higher blocking voltage than a uniform ring spacing at a given total FFR width. Third, comparing (a) to (b) and (c) to (d) we see that in the regime where breakdown voltage is increasing with ring width, a c10-math-0149 ratio of 1 gives a higher blocking voltage at a given total FFR width than a c10-math-0150 ratio of 1.4. Finally, comparing (c) to (d) we conclude that a c10-math-0151 ratio of 1.4 may allow a higher saturation value of blocking voltage than a c10-math-0152 ratio of 1.0.

To summarize, floating field rings can provide blocking voltages of 75–80% of the ideal planar breakdown. They are not sensitive to the activated dopant concentration, and can be fabricated without any additional processing steps. High-voltage SiC devices have been fabricated with as many as 50 concentric rings in their edge termination structure.

10.1.7 Multiple-Floating-Zone (MFZ) JTE and Space-Modulated (SM) JTE

As discussed above, single-zone JTE is sensitive to the activated dopant concentration, and the tight processing tolerance is a challenge in manufacturing. Multi-zone JTE reduces the sensitivity to activated dopant concentration, but brings added processing complexity and cost. The tight processing tolerance of JTE can be avoided with floating field rings, since they are insensitive to the activated dopant concentration. However, it is difficult to identify an algorithm for optimizing FFR terminations.

A fifth termination method combines concepts from both JTE and FFR terminations. Multiple floating zone (MFZ) JTE [11] and space-modulated (SM) JTE [12] are closely-related techniques that achieve the broad dose window of multi-zone JTE with a single implant step. MFZ-JTE employs a series of concentric floating rings having a dose that allows each ring to fully deplete before reaching breakdown. The structure, shown in Figure 10.22, is similar to the FFR structure in Figure 10.19, but the rings are more lightly doped. A careful examination will also reveal that each zone has the same period c10-math-0153, but the c10-math-0154 ratio of successive zones decreases as we move away from the main junction. In this way the effective termination charge in each zone can be tapered smoothly from the full dose at the main junction to zero at the edge of the ring system. Figure 10.23 shows the breakdown voltage on a c10-math-0155 n-type drift layer for single-zone JTE, a 36-zone MFZ-JTE and a 72-zone MFZ-JTE, as a function of ring dose [11]. All three terminations have a total width of c10-math-0156. The MFZ-JTE system provides a much broader range of acceptable doses than the single-zone JTE.

c10f022

Figure 10.22 Cross-section of a c10-math-0157 one-sided step junction protected by multiple-floating-zone JTE termination. The period of each zone is the same, but the c10-math-0158 ratio decreases from the inner to the outer ring.

c10f023

Figure 10.23 Blocking voltage as a function of ring dose for single-zone JTE and two MFZ-JTE terminations on a c10-math-0159 drift region doped c10-math-0160 ( [11] reproduced with permission from IEEE).

Space-modulated (SM) JTE consists of a single broad JTE ring whose outer edge is broken into a number of isolated concentric rings of the same dose, as shown in Figure 10.24. Figure 10.25 shows the breakdown voltage on a c10-math-0161 n-type drift layer for a single-zone JTE, a five-ring SM-JTE with constant c10-math-0162 ratio, and a five-ring SM-JTE with a decreasing c10-math-0163 ratio [12]. Both SM-JTE terminations have a constant ring period c10-math-0164, and all three terminations have a total width of c10-math-0165. The five-ring SM-JTE with decreasing c10-math-0166 ratio provides the broadest range of acceptable doses.

c10f024

Figure 10.24 Cross-section of a c10-math-0167 one-sided step junction protected by a space-modulated JTE termination, consisting of a single JTE ring whose outer edge is split into a number of concentric floating rings.

c10f025

Figure 10.25 Blocking voltage as a function of ring dose for a single-zone JTE and two SM-JTE terminations on a c10-math-0168 drift region doped c10-math-0169 ( [11] reproduced with permission from IEEE). The theoretical planar breakdown voltage for this drift layer is 17.5 kV.

The termination techniques discussed above can be applied with only minor modifications to all the SiC power devices discussed in this book.

10.2 Optimum Design of Unipolar Drift Regions

Unipolar devices such as the Schottky diode, JFET and MOSFET have negligible stored charge, and their switching loss is small compared to their on-state power dissipation. The on-state power dissipation of a unipolar device was given in Equation 7.5 as

where c10-math-0171 is the specific on-resistance and c10-math-0172 is the on-state current density. The unipolar device figure of merit was defined in Equation 7.9 as

where c10-math-0174 is the area of the device, c10-math-0175 the maximum allowable power dissipation, and c10-math-0176 the blocking voltage. The area c10-math-0177 is limited by material, yield, and cost considerations, and the maximum power dissipation c10-math-0178 is limited by the thermal capability of the device and package. The remaining quantity c10-math-0179 is the figure of merit for the device, and it is the goal of the designer to maximize this quantity.

10.2.1 Vertical Drift Regions

All power devices support the terminal voltage in the blocking state by a reverse-biased junction, and for most vertical SiC power devices this is a c10-math-0180 one-sided step junction, such as shown in Figure 7.1. If we neglect field crowding and assume uniform doping in the c10-math-0181 region, the blocking voltage can be written

Here c10-math-0183 is the depletion width at breakdown for a c10-math-0184 one-sided step junction with an infinitely wide c10-math-0185 region, such as shown in Figure 7.1, and c10-math-0186 will be defined below. Assuming a triangular field profile, c10-math-0187 is given by

The critical field EC is given as a function of doping in Equation 10.17 and is plotted in Figure 10.5. We define c10-math-0189 as the effective critical field in a truncated c10-math-0190 junction with a trapezoidal field profile, such as shown in Figure 7.2, and its dependence on doping and width of the drift region is shown in Figure 10.7. The resulting dependence of c10-math-0191 on doping and width of the drift region is given by Equation 10.26 and is plotted in Figure 7.3.

The specific on-resistance of a unipolar device is the sum of all the resistance elements between the terminals, but if the blocking voltage is high the on-resistance is dominated by the resistance of the drift region. The specific on-resistance of the drift region can be written

where c10-math-0193 is the ionized dopant density in the drift region. The mobility parallel to the c-axis in 4H-SiC can be described by the empirical expression [2]

which is plotted in Figure 10.26.

c10f026

Figure 10.26 Electron mobility parallel to the c-axis in 4H-SiC as a function of doping and temperature, as given by Equation 10.29.

For the unipolar drift region, Equations 10.29 establish a relationship between on-resistance and blocking voltage. We can examine this relationship by stepping the doping through a series of values and calculating c10-math-0196 and c10-math-0197 at each doping, with the mobility and critical field re-evaluated at each doping. The on-resistance can then be plotted as a function of blocking voltage, as shown in Figure 10.27. The points on each curve correspond to different values of doping, decreasing from c10-math-0198 on the left to c10-math-0199 on the right (in the sequence 2.0, 1.5, 1.0, 0.7, 0.5, 0.3, 0.2). As the doping is decreased, the depletion region width at breakdown c10-math-0200 increases, following Equation 10.27, and the blocking voltage increases according to Equation 10.26. Eventually the depletion region extends through the entire drift region and the blocking voltage increases more slowly, finally saturating at the value c10-math-0201, as shown by Equation 10.26. The electric field profiles are illustrated in Figure 7.2. Note that even after the blocking voltage saturates, the on-resistance continues to increase as doping is reduced, as shown by Equation 10.28, and the c10-math-0202 characteristic becomes almost vertical.

c10f027

Figure 10.27 Specific on-resistance of an c10-math-0195 drift region in 4H-SiC at room temperature as a function of blocking voltage with drift region width as a parameter.

The optimum design point for the unipolar drift region is the doping – thickness combination that produces the lowest on-resistance at a specified blocking voltage. The dashed line tangent to the curves in Figure 10.27 is the locus of optimum design points and can be described by the empirical equation [2]

10.30 equation

To illustrate the use of these curves, suppose the specified blocking voltage is 3 kV. Then Figure 10.27 tells us the optimum design has a drift region width of c10-math-0204 and a doping of about c10-math-0205 (the doping can be found by counting points on the c10-math-0206 curve from c10-math-0207 on the left to c10-math-0208 at the optimum point, in the sequence 2.0, 1.5, 1.0, etc.). To facilitate identification of the optimum doping, Figure 10.28 plots the figure of merit as a function of doping, with drift region width as a parameter. For a c10-math-0209 drift region, the optimum doping is very close to c10-math-0210. The optimum doping and drift region width for a desired blocking voltage are independent of temperature and are given by the empirical equations [2]

c10f028

Figure 10.28 Unipolar device figure of merit c10-math-0213, evaluated at room temperature as a function of doping with drift region width as a parameter. For a given width, the optimum doping is the doping for which the FOM is a maximum.

For easy reference, Figure 10.29 plots the optimum doping and drift region width from Equations 10.31 and 10.32 as a function of blocking voltage.

c10f029

Figure 10.29 Optimum drift region doping and width as a function of blocking voltage for n-type drift regions in 4H-SiC.

10.2.2 Lateral Drift Regions

When the blocking voltage is not too high, it is feasible to implement power devices using a lateral structure rather than a vertical structure. This places all electrical terminals, the source, gate, and drain, on the top surface. The lateral structure is particularly suited to power integrated circuits, where the power transistor is monolithically integrated with control electronics on the same chip.

Figure 10.30 shows the cross-section of a lateral MOSFET that employs the reduced-surface-field (RESURF) concept [13]. The lightly-doped drain (LDD) extends laterally a distance c10-math-0214 along the surface and has thickness c10-math-0215. In the blocking state, a reverse bias exists between the n-type drain and the grounded p-type body layer underneath, and the LDD is designed so that it completely depletes before the field at the junction reaches the critical field for avalanche breakdown. This requires

10.33 equation

where ND is the LDD doping and c10-math-0217 is the critical field. The significant feature of this structure is that when the LDD is fully depleted, all the field lines from donor charges in the LDD extend vertically, terminating on acceptor charges in the base layer below, as illustrated in Figure 10.31. Since all the donor charges in the LDD are terminated on acceptors in the base layer, x-directed field lines do not terminate on charges in the LDD layer, and there is no field taper in the x-direction, as indicated in the plot. This means that we can increase the drain voltage until c10-math-0218 reaches the critical field c10-math-0219, at which point the blocking voltage is given by

10.34 equation
c10f030

Figure 10.30 Cross-section of a lateral power MOSFET using the reduced-surface-field (RESURF) concept. The p-type base layer is at ground potential.

c10f031

Figure 10.31 Illustration of electric field lines in a lateral RESURF MOSFET in the blocking state. The surface electric field is plotted below the figure, and the dashed line indicates the approximate field when two-dimensional effects are included.

The specific on-resistance of the LDD is

where c10-math-0222 is the resistivity of the LDD layer and c10-math-0223 is the width of the device. Note that the on-resistance depends only on the doping-thickness product NDT and the length of the LDD region.

We can calculate the figure of merit for the RESURF structure, assuming the LDD resistance dominates the device resistance. Using Equations 10.35 we can write

Comparing Equation 10.36 to Equation 7.13, we find that the theoretical limit of a RESURF device is actually four times larger than that of a comparable vertical unipolar power device. However, the situation is not as simple as our analysis would indicate, since we have neglected two-dimensional effects. In practice, field crowding at the ends of the LDD produces field spikes, and the true field is like the dashed line in Figure 10.31. Nevertheless, lateral RESURF devices are a viable alternative to vertical devices in cases where the blocking voltage is not too high.

Although Figures 10.30 and 10.31 illustrate a lateral MOSFET, the same discussion applies to any unipolar device with a lateral drift region, such as a lateral JFET. A modified version of the RESURF principle is also utilized in silicon super-junction MOSFETs, where the drift regions are vertical. To date, no vertical super-junction devices have been reported in SiC.

10.3 Comparison of Device Performance

We conclude this chapter by considering how the different types of power devices may be compared in terms of performance. A fair comparison requires that we evaluate all devices as a function of both blocking voltage and switching frequency, since switching loss is a major limitation on device performance. As discussed in Section 7.1, our figure of merit is the on-state current density c10-math-0225 that each device can carry at a specified switching frequency and blocking voltage, subject to the requirement that the total power dissipation in the device be less than the package limit, which we will arbitrarily take as c10-math-0226. We designate this current density as c10-math-0227, and the device with the highest c10-math-0228 at a given blocking voltage and switching frequency is the preferred device for that application.

The above relationships can be visualized using the construction of Figure 10.32, which compares a power MOSFET and an IGBT in a three-dimensional c10-math-0229 parameter space. The two surfaces represent the loci of constant power dissipation equal to c10-math-0230. Let's first examine how these surfaces depend on blocking voltage and frequency.

c10f032

Figure 10.32 Surfaces of constant power dissipation of a MOSFET and an IGBT in three-dimensional c10-math-0231 parameter space ([15] reproduced with permission from Trans Tech Publications).

To achieve a higher blocking voltage, it is necessary to use a thicker and more lightly-doped drift region. This increases the on-resistance of the device, and hence the on-state power dissipation. To keep the total power below the assumed package limit of c10-math-0232, we must reduce the current density as the blocking voltage is increased. This explains why the surfaces slope downward as c10-math-0233 is increased.

Frequency enters the picture through the switching loss. The switching power dissipation is directly proportional to switching frequency. If we are required to operate at a higher frequency where the switching loss is higher, we must again decrease the current to keep the total power dissipation below c10-math-0234. This explains the slope of the surfaces downward as frequency is increased. The effect is stronger for bipolar devices such as the IGBT, since they dissipate more energy per switching event.

The constant-power surfaces of the MOSFET and the IGBT intersect along a line that can be projected down to the c10-math-0235 plane. This line represents the design points where the two devices have equal performance, in that they permit the same maximum on-state current for operating points along this line. In this example, the IGBT carries higher current at low switching frequencies, while the MOSFET carries higher current at high frequencies. Constructions such as this enable us to visualize the relative performance of different devices in the full current–voltage–frequency parameter space.

The above visualization can be made quantitative as follows. For unipolar devices we can usually assume (i) the I–V characteristics near the origin are linear, and hence can be described by an on-resistance, and (ii) the switching power is small compared to the on-state power dissipation. These assumptions led to Equations 7.8 and 7.9, and the figure of merit for unipolar devices was taken to be c10-math-0236. However, in bipolar devices the current passes through one or more forward-biased pn junctions, and these inject minority carriers that must be removed during the turn-off transient. Hence we must consider the energy dissipated during each switching cycle c10-math-0237. The switching power dissipation is proportional to the switching energy and the frequency, as given in Equation 7.15. Moreover, since the current path in bipolar devices goes through forward-biased diodes whose I–V characteristics are nonlinear, the on-state I–V characteristics often cannot be described by a simple resistance. (This statement depends on the number of forward-biased junctions in the current path. The pin diode, IGBT, and thyristor have an odd number of forward-biased junctions in the current path, and their I–V characteristics are nonlinear near the origin. The BJT has an even number of forward-biased junctions, and its I–V characteristics near the origin are linear.)

To account for both the switching loss and the nonlinear I–V characteristics, we use the procedure described in Section 7.1, which is summarized as follows:

  1. Determine the dependence of switching energy c10-math-0238 on on-state current c10-math-0239 using two-dimensional transient computer simulations. These simulations should include the energy dissipated in the external circuit during switching, and hence a specific load circuit must be assumed.
  2. Determine the dependence of on-state power c10-math-0240 on c10-math-0241 using two-dimensional steady-state computer simulations and obtain a value for c10-math-0242 at the assumed blocking voltage.
  3. For a given switching frequency c10-math-0243, adjust c10-math-0244 using Equation 7.17 until the total power dissipation c10-math-0245 equals c10-math-0246. The resulting current is the desired figure of merit c10-math-0247 for that particular blocking voltage and switching frequency.

This procedure has been used to compare the performance of power MOSFETs and IGBTs at blocking voltages of 15 and 20 kV [14, 15] and will serve as an example of the technique. First, an n-channel DMOSFET and a p-channel DMOS IGBT were individually optimized for the desired blocking voltage, 15 or 20 kV, using computer simulations. The on-state I–V characteristics of the 20 kV devices are shown in Figure 10.33 at two temperatures, 27 and 175 °C. The I–V characteristics of the IGBT are nonlinear, and cannot be represented by an on-resistance. The MOSFET I-V characteristics are linear, but the on-resistance degrades with temperature due to the decrease in mobility with temperature shown in Figure 10.26. In contrast, the IGBT is relatively insensitive to temperature. This is because the ambipolar diffusion length depends on both the diffusion coefficient and the lifetime. As shown in Figure 9.26, the lifetime increases faster with temperature than the diffusion coefficient decreases, and the diffusion length increases slightly with temperature, improving device performance.

c10f033

Figure 10.33 Current–voltage characteristics at 27 °C and 175 °C of an optimized DMOSFET and an optimized p-channel IGBT designed to block 20 kV ([15] reproduced with permission from Trans Tech Publications).

The on-state power c10-math-0248 or c10-math-0249 can be calculated at any c10-math-0250 directly from Figure 10.33, and is plotted as a function of current in Figure 10.34. The switching energy is obtained as a function of on-current from transient simulations using a clamped inductive load (modeled by a current source of magnitude c10-math-0251), and the resulting c10-math-0252 is plotted as a function of current in Figure 10.35. These plots are used in conjunction with Equation 7.17 to determine the c10-math-0253 that corresponds to a total power dissipation of c10-math-0254. This is the desired figure of merit c10-math-0255. Figure 10.36 shows c10-math-0256 for each device as a function of switching frequency. We observe that at low switching frequencies, the IGBT provides higher current than the MOSFET. This is due to conductivity modulation of the drift region, which reduces the forward voltage drop at high currents, as seen in Figure 10.33. However, the minority carriers that provide the conductivity modulation must be removed during each turn-off event, and the reverse current transient gives rise to a power dissipation that is proportional to frequency. As frequency is raised, the on-current c10-math-0257 must be reduced to keep the total power dissipation below c10-math-0258, and at higher switching frequencies the on-current of the IGBT falls below that of the MOSFET. Thus the MOSFET is the preferred device for high-frequency applications, while the IGBT is superior at low frequencies.

c10f034

Figure 10.34 On-state power dissipation of the DMOSFET and IGBT of Figure 10.33 as a function of on-state current density, (a) at 27 °C and (b) at 175 °C ( [14] reproduced with permission from Tomohiro Tamaki).

c10f035

Figure 10.35 Switching energy of the DMOSFET and IGBT of Figure 10.33 as a function of on-state current density, as obtained from two-dimensional transient simulations using a clamped inductive load, (a) at 27 °C and (b) at 175 °C ( [14] reproduced with permission from Tomohiro Tamaki).

c10f036

Figure 10.36 Current density of the DMOSFET and IGBT of Figure 10.33 assuming a total power dissipation of c10-math-0259, plotted as a function of switching frequency, (a) at 27 °C and (b) at 175 °C ( [13] reproduced with permission from Tomohiro Tamaki).

The methodology above can be used to compare bipolar or unipolar devices for any desired blocking voltage and switching frequency. In general, bipolar devices perform better at high blocking voltages and low frequencies, while unipolar devices are superior at low blocking voltages and high frequencies. However, at sufficiently low blocking voltages, MOSFETs and JFETs are superior to IGBTs and thyristors, regardless of the frequency. Why is this? There are two reasons. First, at low blocking voltages the drift region resistance is relatively small, and does not need to be reduced by conductivity modulation. Second, IGBTs and thyristors have an odd number of pn junctions in their current paths. When all junctions are forward biased, a net voltage drop of one forward-biased diode appears in series with the remainder of the device. This adds an additional component to the static power dissipation of the IGBT and thyristor. (In contrast, the BJT has an even number of junctions in its current path, and the voltage drops of two oppositely-directed junctions cancel out.)

At sufficiently high frequencies, unipolar devices are superior at all blocking voltages, as seen in the comparison of pin and Schottky diodes in Figure 7.9. This is because the switching loss of bipolar devices is proportional to frequency and becomes the dominant loss at high frequencies, forcing a continuous reduction in c10-math-0260 to keep c10-math-0261 below c10-math-0262.

In all comparisons involving power devices, it is important to consider both high and low temperature extremes, since device performance is sensitive to temperature. A device dissipating c10-math-0263 will have a junction temperature well above ambient. For this reason, analyses performed only at room temperature provide an incomplete picture of device performance.

References

  1. [1] Konstantinov, A.O., Wahab, Q., Nordell, N. and Lindefelt, U. (1997) Ionization rates and critical fields in 4H silicon carbide. Appl. Phys. Lett., 71 (1), 90–92.
  2. [2] Morisette, D.T. (2001) Development of robust power Schottky barrier diodes in silicon carbide. PhD thesis. Purdue University.
  3. [3] Loh, W.S., David, J.P.R., Ng, B.K. et al. (2009) Temperature dependence of hole impact ionization coefficient in 4H-SiC photodiodes. Mater. Sci. Forum, 615–617, 311–314.
  4. [4] Konstantinov, A.O., Nordell, N., Wahab, Q. and Lindefelt, U. (1998) Temperature dependence of avalanche breakdown for epitaxial diodes in 4H-silicon carbide. Appl. Phys. Lett., 73 (13), 1850–1852.
  5. [5] Davies, R.L. and Gentry, F.E. (1964) Control of electric field at the surface of p-n junctions. IEEE Trans. Electron. Devices, 11 (7), 313–323.
  6. [6] Adler, M.S. and Temple, V.K. (1978) Maximum surface and bulk electric fields at breakdown for planar and beveled devices. IEEE Trans. Electron. Devices, 25 (10), 1266–1270.
  7. [7] Temple, V.A.K. (1977) Junction termination extension (JTE): a new technique for increasing avalanche breakdown voltage and controlling surface electric fields in p-n junctions. International Electron Devices Meeting Technical Digest, pp. 423–426.
  8. [8] Khan, I.A. (2002) High voltage SiC MOSFETs. PhD thesis. Purdue University.
  9. [9] Wang, X. and Cooper, J.A. (2004) Optimization of JTE edge termination for 10 kV power devices in 4H-SiC. Mater. Sci. Forum, 457–460, 1257–1262.
  10. [10] Singh, S., Tjandra, E. and Cooper, J. A. (2003) High-voltage termination of SiC devices using floating field rings. unpublished, reproduced with permission from James A. Cooper.
  11. [11] Sung, W., Van Brunt, E., Baliga, B.J. and Huang, A.Q. (2011) A new edge termination technique for high-voltage devices in 4H-SiC-multiple-floating-zone junction termination extension. IEEE Electron. Device Lett., 32 (7), 880–882.
  12. [12] Feng, G., Suda, J. and Kimoto, T. (2012) Space-modulated junction termination extension for ultra-high voltage p-i-n diodes in 4H-SiC. IEEE Trans. Electron. Devices, 59 (2), 414–418.
  13. [13] Appels, J.A. and Vaes, H.M.J. (1979) High voltage thin layer devices (RESURF) devices. IEEE International Electron Devices Meeting Technical Digest, pp. 238–241.
  14. [14] Tamaki, T., Walden, G. G., Sui, Y., and Cooper, J. A. (2007) On-state and switching performance of high-voltage 4H-SiC DMOSFETs and IGBTs. International Conference on Silicon Carbide and Related Materials, Otsu, Japan.
  15. [15] Tamaki, T., Walden, G.G., Sui, Y. and Cooper, J.A. (2009) On-state and switching performance of high-voltage 4H-SiC DMOSFETs and IGBTs. Mater. Sci. Forum, 600–603, 1143–1146.
..................Content has been hidden....................

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