Chapter 9
Bipolar Power Switching Devices

9.1 Bipolar Junction Transistors (BJTs)

Bipolar transistors were the first power transistors developed in silicon, but they have now been largely displaced by silicon insulated-gate bipolar transistors (IGBTs) and thyristors. This occurred because silicon bipolar junction transistors (BJTs) suffer from a phenomenon known as “second breakdown” that limits the safe operating area (SOA) of the device. As will be shown below, the higher critical field for avalanche breakdown in SiC essentially eliminates second breakdown, making high-performance power BJTs practical. BJTs are particularly attractive for high-temperature applications, since they are not dependent on a gate oxide for their operation, and are not subject to the oxide reliability limitations of metal-oxide-semiconductor field effect transistors (MOSFETs) and IGBTs.

As is our custom when considering a new device, we begin our discussion with a review of the basics of BJT operation. We then consider a number of special effects that occur during high-current operation of power BJTs. Figure 9.1 shows a basic c09-math-0001 BJT, along with standard current and voltage definitions. In these devices the emitter is more heavily doped than the base, and the base is more heavily doped than the collector, so c09-math-0002. We can identify four internal junction currents, the hole and electron currents crossing the emitter–base (EB) junction c09-math-0003 and the hole and electron currents crossing the collector–base (CB) junction c09-math-0004. The polarities in the figure correspond to positive current flow, and we should remember that electron flux is opposite to direction of electron current. Voltages c09-math-0005 and c09-math-0006 are developed across the junction depletion regions, shown cross-hatched in the figure. The widths of the emitter, base, and collector, measured from the metallurgical junctions, are c09-math-0007, and c09-math-0008, and the width of the neutral portion of the base, measured between the edges of the depletion regions, is simply c09-math-0009.

c09f001

Figure 9.1 Basic structure of an c09-math-0029 BJT showing internal current components c09-math-0030, and c09-math-0031, terminal currents c09-math-0032, and c09-math-0033, and c09-math-0034, and c09-math-0035 coordinate systems. The arrows indicate assumed directions for positive current.

9.1.1 Internal Currents

Our first goal is to obtain equations for the four internal currents c09-math-0010, and c09-math-0011 in terms of the device parameters and the terminal voltages c09-math-0012 and c09-math-0013. We can then write expressions for the terminal currents c09-math-0014, and c09-math-0015, and from these we will obtain equations for important performance parameters such as current gain, specific on-resistance, and so on. We initially assume low-level injection in all regions, since this allows us to use the minority carrier diffusion equations (MCDEs) to obtain the desired internal currents. We will later remove this restriction in the base and collector.

The solution to the MCDE in the emitter and collector proceeds in the same way as in the neutral regions of a pn diode discussed in Section 7.3. We first write the general solution of the MCDE in the emitter and collector, subject to two boundary conditions in each region. One boundary condition, the carrier density at the edge of the depletion region, is provided by the law of the junction, Equation 7.25. The second boundary condition is based on the assumption that the minority carrier densities far from the junctions remain at their equilibrium values. However, in SiC BJTs it may not be accurate to assume that the emitter ohmic contact is infinitely far from the junction. In this case, we must set the minority carrier density at the position of the ohmic contact to its equilibrium value. The solution for c09-math-0016 in the emitter involves hyperbolic functions (see Appendix B), and can be written

Here c09-math-0018 is the intrinsic carrier concentration in the emitter, c09-math-0019 is the ionized dopant concentration in the emitter, and c09-math-0020 is the hole diffusion length in the emitter, where c09-math-0021 is the hole diffusion coefficient and c09-math-0022 is the hole lifetime. We specify c09-math-0023 in the emitter separately from c09-math-0024 in the other portions of the device, since the emitter is usually so heavily doped that bandgap narrowing occurs, resulting in a higher intrinsic carrier concentration. The hole current crossing the EB junction is obtained from the derivative of Equation 9.1 at the depletion edge x = 0,

where c09-math-0026 is the area of the junction. In the case of a “long” emitter where c09-math-0027, Equation 9.2 reduces to

The solution in the collector is similar to that in the emitter. By analogy with Equation 9.2 we can write the hole current crossing the CB junction as

which in the case of a “long” collector reduces to

Equations 9.3 and 9.5 are similar in form to the Shockley diode Equation 7.26.

The solution for the minority carrier density in the base is obtained in the same manner – we find the general solution of the MCDE, subject to the boundary conditions at the EB and CB junctions provided by the law of the junction. In a typical SiC BJT we cannot always assume c09-math-0038, and the full solution for c09-math-0039 will involve hyperbolic functions, namely

The electron current crossing the EB junction is obtained from the derivative of Equation 9.6 evaluated at c09-math-0041, and the electron current crossing the CB junction is obtained from the derivative evaluated at c09-math-0042. After some algebra, we can write

and

If the base width is much shorter than the minority carrier diffusion length in the base, we can set c09-math-0045, and the above equations reduce to

9.9 equation

Equations 9.7, and 9.8 are the desired equations for the four internal currents in the BJT of Figure 9.1.

9.1.2 Gain Parameters

We can now write equations for some of the important performance parameters of the BJT. The common-base current gain c09-math-0047 is defined as the ratio of collector current to emitter current. We assume the BJT is operated in the forward-active mode with the EB junction forward biased and the CB junction reverse biased, so that c09-math-0048 and c09-math-0049. Under forward-active biasing the c09-math-0050 terms in the above equations are small compared to the c09-math-0051 terms, and the c09-math-0052 terms are small compared to the c09-math-0053 terms. The current component c09-math-0054 represents the reverse-bias leakage current of the CB junction, and from Equation 9.4 we see it is very small and can be neglected. Therefore, the common-base current gain can be written

where c09-math-0056 is defined as the base transport factor and c09-math-0057 is the emitter injection efficiency. The base transport factor is the fraction of electrons injected from the emitter that diffuse across the base and are swept across the collector junction. The emitter injection efficiency is the fraction of current crossing the EB junction that is due to electrons injected from the emitter into the base. The base transport factor is obtained by dividing Equation 9.8 by Equation 9.7. Under forward-active biasing, c09-math-0058 can be simply written

Likewise, using Equations 9.7 and 9.2, the emitter injection efficiency is

Inserting Equations 9.11 and 9.12 into Equation 9.10, we can write the common-base current gain as

The common-emitter current gain c09-math-0062 is the ratio of collector current to base current,

Inserting Equation 9.13 into Equation 9.14, we can write

9.15 equation

We can simplify the above equations under certain conditions. If the base is narrow compared to a diffusion length, the hyperbolic functions involving c09-math-0065 can be replaced by the first term of their Taylor series expansions, namely,

Moreover, if the emitter is long compared to a diffusion length, the hyperbolic functions involving c09-math-0067 can be simplified using

Under these conditions, c09-math-0069 and we can write

From this, it follows that

If the emitter is short compared to a diffusion length, c09-math-0072, we simply replace c09-math-0073 by c09-math-0074 in the above equations. Although these simplifications are appealing, a word of caution is in order. The “narrow base” and “long emitter” assumptions are often invoked for silicon BJTs, but they may not be accurate for all SiC power BJTs. When in doubt, it is advisable to use the full expressions that involve hyperbolic functions.

9.1.3 Terminal Currents

We next wish to obtain expressions for the terminal currents c09-math-0075, and c09-math-0076. This is easily done by recognizing that c09-math-0077, and c09-math-0078. Adding Equations 9.7 and 9.2 we obtain

Similarly, adding Equations 9.8 and 9.4 yields

The equation for c09-math-0081 follows by subtracting Equation 9.21 from Equation 9.20.

The above equations initially seem formidable, but a closer analysis reveals symmetries that lead to a simpler formulation. We note that each equation is the algebraic sum of two Shockley diode equations of the form given in Equation 7.26. We can exploit this symmetry by writing Equations 9.20 and 9.21 in the alternate forms

and

Equations 9.22 and 9.23 are known as the Ebers–Moll equations for the BJT. The four constants in the Ebers–Moll equations can be determined by comparing Equations 9.22 and 9.23 to Equations 9.20 and 9.21. For convenience, we summarize these parameters below.

As before, the approximations in Equations 9.16 and 9.17 can be applied to the above equations, if justified in a particular situation.

Equations 9.24 allow us to calculate all four constants in the Ebers–Moll equations for the BJT terminal currents. The Ebers–Moll equations are associated with a simple and easily-remembered equivalent circuit for the BJT known as the Ebers–Moll model, shown in Figure 9.2. The Ebers–Moll equations and model are valid in all biasing regimes of the BJT: forward-active, saturation, inverse-active, and cutoff, and they form the basis for the BJT model used in popular circuit analysis programs such as c09-math-0085.

c09f002

Figure 9.2 The Ebers–Moll equivalent circuit model of the npn BJT. This model can be used in all operating regimes: forward-active, saturation, inverse-active, and cutoff.

Consider the Ebers–Moll model for a BJT in the forward-active mode, with the EB junction forward biased and the CB junction reverse biased. With the CB junction reverse biased, the diode current c09-math-0086 is just the reverse leakage current −c09-math-0087, which is small and can be neglected. Likewise, the dependent current generator c09-math-0088 c09-math-0089 can be neglected. Diode c09-math-0090 then represents the hole and electron currents crossing the EB junction under forward bias, and the dependent current generator c09-math-0091 c09-math-0092 represents the fraction of electron current injected from the emitter that reaches the collector (keep in mind that the flow of electrons is opposite to the positive direction of electron current). If the BJT were operated in the inverse-active mode with the EB junction reverse biased and the CB junction forward biased, the roles of the c09-math-0093 and c09-math-0094 diodes would be reversed: the collector would inject electrons into the base, and the emitter would sweep them out. In saturation, both junctions are forward biased and all four elements of the Ebers–Moll model are active.

9.1.4 Current–Voltage Relationship

The Ebers–Moll model can be used to obtain a single equation for the c09-math-0095 characteristics with base current as a parameter. We first write c09-math-0096 in terms of the Ebers–Moll circuit model as

where, as illustrated in Figure 9.2,

and

We can eliminate VBC in Equation 9.27 by writing c09-math-0100, resulting in

Inserting Equations 9.26 and 9.28 into Equation 9.25 then provides an equation for c09-math-0102 in terms of c09-math-0103 and c09-math-0104,

Equation 9.29 can be solved for c09-math-0106, yielding

From the Ebers–Moll model we can write the collector current as

We now substitute Equation 9.30 into Equations 9.26 and 9.28 to eliminate c09-math-0109, then insert Equations 9.26 and 9.28 into Equation 9.31 to obtain the desired equation for c09-math-0110 as a function of c09-math-0111 with c09-math-0112 as a parameter. After some algebra we can write

Equation 9.32 is valid in all biasing regimes: forward-active, saturation, inverse-active, and cut-off. Despite its apparent complexity, Equation 9.32 is a simple algebraic equation involving four constants (the four Ebers–Moll parameters) and the variables c09-math-0114 and c09-math-0115. Equation 9.32 can be used to generate a plot of c09-math-0116 as a function of c09-math-0117 for various values of base current c09-math-0118, and such a plot is shown in Figure 9.3. The parameters used to generate this plot are representative of a typical 4H-SiC c09-math-0119 BJT.

c09f003

Figure 9.3 Current–voltage characteristics of an c09-math-0120 BJT in 4H-SiC at room temperature, calculated using Equation 9.32. In this example, c09-math-0121, and c09-math-0122, and c09-math-0123 are 1 ns, 10 ns, and c09-math-0124, and c09-math-0125, and c09-math-0126 are 78, 123, 565, and c09-math-0127, respectively. The common-emitter current gain is 25 in the forward-active region and 1.1 in the inverse-active region.

The equations developed to this point involve a number of assumptions that may not be valid in all regions of operation. This is especially true of power BJTs, since they operate at high current densities where our assumptions of low-level injection fail, first in the collector and then in the base. Other effects occur at high voltages, and still others at high temperatures. In addition, there are important lateral effects that are not captured in our one-dimensional analysis. In the following sections we will discuss the most important of these effects, and how they influence practical SiC power BJTs.

9.1.5 High-Current Effects in the Collector: Saturation and Quasi-Saturation

To support a high blocking voltage in the off state, the power BJT incorporates a thick, lightly-doped collector drift region between the CB junction and the c09-math-0128 substrate, as shown in Figure 9.4. As with the power junction field-effect transistor (JFET) and power MOSFET, this drift region is designed to support the desired blocking voltage, with doping and thickness given by Equations 7.10 and 7.11. In the absence of conductivity modulation, the drift region would add a series resistance given by Equation 7.12 to the basic BJT described above, but in the saturation regime the drift region is partially or totally conductivity modulated. The actual resistance and associated voltage drop added by the collector drift region will be considered next.

c09f004

Figure 9.4 Implementation of a practical power BJT. All layers are epitaxially grown to avoid the reduced lifetimes associated with implanted regions.

Figure 9.5 illustrates the c09-math-0130 relationships of an c09-math-0131 power BJT with a thick lightly-doped collector drift region, and Figure 9.6 shows the corresponding minority carrier densities in the device at operating points (a), (c), and (e). The behavior in saturation consists of two distinct regimes: saturation and quasi-saturation. This can be understood as follows. In the forward-active region (a), the CB junction is reverse biased and minority carriers are not injected into the collector drift region. The drift region is not conductivity modulated, and the carrier densities are illustrated in Figure 9.6a. As c09-math-0132 is reduced, the reverse bias on the CB junction is reduced and c09-math-0133 eventually reaches zero at point (b) in Figure 9.5. This is the boundary between the forward-active and saturation regimes. As c09-math-0134 is reduced further, the CB junction becomes forward biased, injecting holes into the c09-math-0135 collector drift region. Since the collector is lightly doped, even a small injection of holes places the region near the junction into high-level injection, and the minority carrier density at bias point (c) is shown in Figure 9.6c. Here the drift region is conductivity modulated out to a distance c09-math-0136, but not beyond. The carrier densities vary linearly with distance from c09-math-0137 to c09-math-0138, as will be explained below. As we continue to reduce c09-math-0139, the CB junction becomes even more forward biased, increasing the injection of holes into the drift region, and at bias point (d) the modulated portion extends across the entire drift region, that is, c09-math-0140. The resistance of the drift region has now reached a very low value, and the c09-math-0141 characteristic in Figure 9.5 follows a steep slope toward the origin. Thus, the term “saturation” in the power BJT refers to the condition where the drift region is fully conductivity-modulated, and the term “quasi-saturation” refers to the condition where the drift region is partially conductivity-modulated.

c09f005

Figure 9.5 Illustration of the c09-math-0129 characteristics of a power BJT with a thick drift region.

c09f006

Figure 9.6 Minority carrier densities in the neutral regions of a power BJT in (a) forward-active, (c) quasi-saturation, and (e) saturation biasing regimes (see points a, c and e in Figure 9.5).

We will now develop equations describing the carrier density in the drift region and the voltage drop across the drift region in saturation and quasi-saturation. Since the region must remain charge neutral, we can write

9.33 equation

The hole concentration at c09-math-0143 is given by the law of the junction, Equation 7.25,

Since the drift region is lightly doped, c09-math-0145 can exceed c09-math-0146 at even moderate forward biases c09-math-0147, so high-level injection prevails, and in this region we can write

The electron and hole currents in the drift region are

We now assert that the injected hole current flowing into the drift region from the base, c09-math-0151, is small compared to the electron current flowing across the base from the emitter, c09-math-0152. The situation in the BJT is obviously different from the pin diode, where electrons are injected from the c09-math-0153 region and holes from the c09-math-0154 region. In the BJT drift region, the primary electron current comes from electrons injected from the emitter c09-math-0155 that diffuse across the base into the collector. The hole current injected from the base into the collector is much smaller, due to the gain of the transistor. Recall that c09-math-0156. Assuming c09-math-0157, the base current is much smaller than the collector current. The base current in saturation consists of holes injected into the emitter c09-math-0158 plus holes injected into the collector c09-math-0159 plus holes recombining in the base, so we can be sure that c09-math-0160. Thus the collector current is almost totally due to electrons, and we can set c09-math-0161 in Equation 9.37. Solving for the electric field and employing the Einstein relationship, we find that

The collector current, which is due primarily to electrons, can be obtained by combining Equation 9.36 with Equations 9.35, and 9.38,

Again using the Einstein relationship, Equation 9.39 can be written

This equation can be solved by cross-multiplying both sides by c09-math-0165 and integrating with respect to c09-math-0166. Performing the integration and solving for c09-math-0167 yields

9.41 equation

Since c09-math-0169 in saturation, the third term can be dropped, making the hole density a linearly decreasing function of position,

The reader should recognize that the transport parameters c09-math-0171 and c09-math-0172 in Equation 9.39 and the following equations are those of majority carriers (electrons) in the n-type collector drift region. This is different from our earlier development in Equations 9.32, where the transport parameters in each region were those of minority carriers.

As an aside, we note that Equation 9.42 could have been deduced directly from the ambipolar diffusion equation (ADE) Equation 7.39. If recombination in the drift region is negligible, we can set c09-math-0173. In steady-state c09-math-0174, so the ADE reduces to simply c09-math-0175. This means c09-math-0176 is a linear function of position, and we can write

Since c09-math-0178, Equation 9.40 can be rewritten in the form

Inserting Equation 9.44 into Equation 9.43 leads directly to Equation 9.42.

Having obtained an expression for the electron and hole densities in the collector drift region as a function of current, Equation 9.42, we now wish to calculate the voltage drop across the drift region. This voltage drop can then be added to the c09-math-0180 in Equation 9.32 to obtain an c09-math-0181 relation for the power BJT. To begin, we note from Equation 9.42 that the c09-math-0182 value at which the hole density equals the background doping density is

9.45 equation

Inserting Equation 9.34 for c09-math-0184 yields

Here, c09-math-0186 refers to the voltage drop across the internal CB junction, which is less than the voltage between the collector and base terminals because of the voltage developed across the collector drift region and the substrate. In Equation 9.46, c09-math-0187 is obtained from the internal c09-math-0188 by setting c09-math-0189, with c09-math-0190 given by Equation 9.30.

We assume the drift region is conductivity modulated over the region c09-math-0191 where the hole density exceeds the background doping density, but not in the region c09-math-0192. The voltage drop c09-math-0193 across the unmodulated portion of the drift region is then given by

The voltage drop across the conductivity-modulated portion c09-math-0195 is obtained by integrating the electric field given by Equation 9.38 over the region c09-math-0196. Thus we can write

where c09-math-0198 is given by Equation 9.34. In evaluating Equation 9.34, we set c09-math-0199, with c09-math-0200 given by Equation 9.30. For SiC power BJTs, c09-math-0201 is typically less than c09-math-0202 and can be neglected compared to other voltage drops in the device.

Having obtained expressions for the voltage drop across the collector drift region, we are now in a position to modify the c09-math-0203 relation given in Equation 9.32 to include these effects. The procedure is as follows. We chose a value for base current c09-math-0204 and step the internal c09-math-0205 to calculate c09-math-0206 using Equation 9.32. To obtain the terminal c09-math-0207, we add Equations 9.47 and 9.48 to the internal c09-math-0208 using Equation 9.46 for c09-math-0209. Figure 9.7 shows c09-math-0210 curves for the transistor of Figure 9.3 with the drift region voltage included (solid lines). For reference, the dotted lines show the curves of Figure 9.3 that do not include the drift region voltage. Points (a) and (b) on the c09-math-0211 curve denote boundaries between different modulation modes. From the origin to point (a), the drift region is fully conductivity modulated c09-math-0212 and its differential resistance is very small. Between (a) and (b), the drift region is partially conductivity modulated c09-math-0213 and the differential resistance is higher. To the right of point (b), the injection level c09-math-0214 is insufficient to produce any conductivity modulation c09-math-0215, and the differential resistance is that of the full unmodulated drift region.

c09f007

Figure 9.7 c09-math-0216 characteristics including the voltage drop across the collector drift region in the BJT of Figure 9.3. The dotted curves are the characteristics of the BJT given by the basic Ebers–Moll model of Figure 9.3, without the voltage drop across the collector drift region.

Characteristics like those in Figure 9.7 are often not seen in experimental devices, for several reasons. First, at the somewhat higher dopings needed to maximize c09-math-0220, the drift region often never reaches full conductivity modulation as c09-math-0221, and point (a) lies very close to the origin. Figure 9.8 shows the device of Figure 9.7 with a drift region doping of c09-math-0222 instead of c09-math-0223. For the c09-math-0224 drift region used here, this doping provides the optimum c09-math-0225 and a theoretical blocking voltage of 1580 V. At this higher doping, the device remains in quasi-saturation almost to the origin. A second reason is that lateral effects such as current spreading in the drift region and the voltage drop due to spreading resistance in the base often obscure the breakpoints in the idealized situation of Figure 9.7. Finally, we should caution that the development leading to Equation 9.32 assumes low-level injection in all regions, whereas in quasi-saturation and saturation the collector drift region is in high-level injection. This will alter the parameters c09-math-0226 and c09-math-0227 in Equation 9.24 and subsequent equations, and this effect has not been included in Equation 9.32 or Figures 9.7 and 9.8.

c09f008

Figure 9.8 c09-math-0217 of the BJT of Figure 9.7 with collector doping of c09-math-0218 instead of c09-math-0219. The higher doping prevents complete modulation of the drift region, leaving the device in quasi-saturation until very near the origin.

9.1.6 High-Current Effects in the Base: the Rittner Effect

We have discussed high-level injection in the collector and examined how conductivity modulation can reduce the drift region voltage drop in saturation. We now turn to high-level injection in the base. When the BJT is operated at high currents, the injection of electrons from the emitter into the base can become large enough that the electron density in the base exceeds the ionized dopant density. When this occurs, charge neutrality demands that the hole concentration in the base also increase so that c09-math-0228. The higher hole concentration increases the back injection of holes from the base into the emitter. This is intuitively reasonable, since if we have more holes per unit volume in the base, we expect a greater flow of holes from base to emitter under forward bias conditions. The increased hole flow must be supplied by an increased base current, and this means a lower current gain, since c09-math-0229. This phenomenon is known as the Rittner Effect. We will now develop equations to describe this effect, culminating in an expression for c09-math-0230 as a function of collector current.

To solve for c09-math-0231, we need to obtain expressions for the collector and base currents under conditions of high-level injection in the base. If we neglect recombination in the base, the base current in the forward-active region consists exclusively of holes injected from the base into the emitter, the component identified as c09-math-0232 in Figure 9.1. Since the emitter remains in low-level injection, this current can be found by solving the MCDE in the emitter, subject to the boundary condition imposed by the hole density at the depletion edge, c09-math-0233. Our first task is to find this boundary condition.

Invoking the law of the junction on the emitter side of the EB depletion region, we can write

9.49 equation

Likewise, on the base side of the EB depletion region, we can write

9.50 equation

In these equations, c09-math-0236 denotes the edge of the depletion region in the emitter and 0 denotes the edge of the depletion region in the base. The above equations can be combined to give the hole density in the emitter at the depletion edge,

Since the emitter remains in low-level injection, the electron density c09-math-0238 remains at its equilibrium value c09-math-0239. However, if the base is in high-level injection, the hole density c09-math-0240 will be greater than the doping density in the base. Charge neutrality in the base requires that

9.52 equation

Solving for c09-math-0242, we can write

Evaluating Equation 9.53 at c09-math-0244 and inserting into Equation 9.51 yields

We see that the boundary condition for holes in the emitter c09-math-0246 now depends on the injection level in the base c09-math-0247. From the law of the junction, we can write

Inserting Equation 9.55 into Equation 9.54 yields

Having obtained the boundary condition for the hole density in the emitter, we are now in a position to write expressions for the base and collector currents, and from their ratio we can find c09-math-0250. Assuming the emitter width c09-math-0251 is long compared to the minority diffusion length c09-math-0252, the solution of the MCDE for holes in the emitter is a decaying exponential,

9.57 equation

The hole current flowing into the neutral emitter at c09-math-0254 is given by

As stated earlier, with negligible recombination in the base, c09-math-0256. The collector current c09-math-0257 is the electron current flowing across the base from the emitter, namely, the components c09-math-0258 in Figure 9.1 (recall that electron particle flow is in the opposite direction to electron current). With negligible recombination in the base, the electron density decreases linearly from the emitter edge to the collector edge (see Figure 9.6a), and the collector current can be written

This gives us another expression for c09-math-0260, namely

We now divide Equation 9.59 by Equation 9.58 to obtain

9.61 equation

Inserting Equation 9.55 for c09-math-0263 and Equation 9.56 for c09-math-0264, and substituting Equation 9.60 into Equation 9.56 yields

The last factor in Equation 9.62 describes the decrease in c09-math-0266 due to high-level injection in the base. As the collector current density c09-math-0267 is reduced, Equation 9.62 approaches the low-injection c09-math-0268 given in Equation 9.19, and we can rewrite Equation 9.62 in the form

9.63 equation

where c09-math-0270 is the Rittner current density given by

9.64 equation

The Rittner current can be regarded as the collector current density where c09-math-0272 has fallen to half of its low-injection value. It is obviously desirable to have a high value of the Rittner current, which suggests increasing the base doping and/or reducing the base width. However, c09-math-0273 given in Equation 9.19 is inversely proportional to c09-math-0274, so increasing the doping has the undesirable effect of reducing c09-math-0275. Reducing base width increases both c09-math-0276 and c09-math-0277, but to avoid punch-through in the base, the doping-thickness product needs to obey

Taking the equality in Equation 9.65, the optimum c09-math-0279 and c09-math-0280 can be written in terms of the critical field as

9.66 equation

and

Thus, for highest gain and highest Rittner current, we should reduce c09-math-0283 while increasing c09-math-0284 to satisfy the equality in Equation 9.65, thereby preventing punch-through. Equation 9.67 indicates another advantage of SiC for power BJTs, namely a higher Rittner current due to the higher critical field, but the Rittner current is reduced somewhat at room temperature by the incomplete ionization of acceptor dopants in the base.

9.1.7 High-Current Effects in the Collector: Second Breakdown and the Kirk Effect

In the forward-active region of operation, the c09-math-0285 emitter injects electrons into the base, where they diffuse to the reverse-biased CB junction and are swept into the c09-math-0286 collector drift region. At high current densities, typical of power BJTs, a significant density of electrons exists in the base, throughout the CB depletion region, and in the collector drift region. As discussed in the previous section, when the density of electrons in the base exceeds the base doping, the base enters high-level injection, and this leads to a reduction in current gain c09-math-0287 through the Rittner Effect. We now wish to consider conditions in the CB depletion region and the c09-math-0288 collector drift region.

Focusing first on the CB depletion region, the electric field at any point within the depletion region must obey Poisson's equation. On the collector side of the metallurgical junction we may write

where c09-math-0290 is the density of electrons drifting across the depletion region, and we have redefined our c09-math-0291 coordinate system to place the origin at the CB metallurgical junction. c09-math-0292 here represents the total donor density, since donors atoms in the depletion region are fully ionized. The CB junction is strongly reverse biased and the electric field is high, so we may assume that electrons are moving at their saturated drift velocity c09-math-0293. Under this assumption, the electron density in the depletion region is uniform with respect to position, and we can write

We can calculate the electric field by inserting Equation 9.69 into Equation 9.68 and integrating with respect to c09-math-0295, resulting in

In equilibrium the collector current c09-math-0297 is zero, and Equation 9.69 shows there are no electrons in the depletion region. The electric field in this case is illustrated by curve (a) in Figure 9.9. The depletion region spreads much further into the c09-math-0298 collector due to the asymmetry in doping between the base and collector. As the collector current is increased, the electron density given by Equation 9.69 becomes significant compared to the (very low) doping density c09-math-0299, and the field taper given by Equation 9.68 is reduced, resulting in profile (b) in Figure 9.9. Here the depletion region extends through the entire collector drift region, and the field falls rapidly inside the c09-math-0300 substrate. If the current is increased even more, the electron density given by Equation 9.69 will eventually equal the doping density c09-math-0301, at which point the slope of the field profile becomes zero, curve (c). At even higher currents, the electron density will exceed the doping density and the net charge density changes sign. The electric field now increases with distance, resulting in profile (d). Here the peak field has shifted from the CB junction to the c09-math-0302 junction. If the current continues to increase, the field at the CB junction eventually goes to zero, profile (e). The current density corresponding to profile (e) is called the Kirk current, and can be calculated from Equation 9.70 by setting c09-math-0303, and c09-math-0304, resulting in

c09f009

Figure 9.9 Electric field profiles in the collector when the BJT is carrying a high current density in the forward-active region. The device cross-section illustrates only the emitter, base, and collector doping regions, and does not distinguish between depletion and neutral regions.

If c09-math-0306 exceeds the Kirk current c09-math-0307, the electric field shifts toward profile (f). When the peak field at the c09-math-0308 junction reaches the critical field for avalanche breakdown, as illustrated by profile (f), the device enters second breakdown. Holes generated by impact ionization in the high-field region at the c09-math-0309 junction flow toward the base and are injected into the emitter, acting as additional base current. This results in a much larger electron current from the emitter, further increasing the collector current and providing positive feedback to the breakdown process. It is important to note that this can occur at collector voltages below that required to cause avalanche breakdown at zero current. This is because at high current densities the field taper in the collector drift region given by Equation 9.70 and shown as profile (f) can exceed the normal field taper when the current is zero, profile (a). Since the area under the electric field remains equal to c09-math-0310, which is held constant, the peak field in (f) can be higher than the peak field in (a), even though the voltage has not increased.

Now let us see how these concepts can be related to the critical field of the semiconductor. When the device is in the blocking state with c09-math-0311, the electric field is shown in profile (a). As discussed in Section 7.1.1, in a non-punch-through (or NPT) design, the drift layer is totally depleted when the peak field just equals the critical field. Gauss' law and Poisson's equation lead to the relations

9.72 equation

and

9.73 equation

Inserting these expressions into Equation 9.71 yields

9.74 equation

This indicates that for a given blocking voltage c09-math-0315, the Kirk current increases as the square of the critical field. The higher critical field of SiC leads to a much higher value for the Kirk current, effectively eliminating second breakdown as an issue in SiC BJTs.

We note that when the electric field has the form of profile (f), the field is zero over the portion of the collector drift region nearest the CB metallurgical junction. Poisson's equation requires that when the field is uniform (and in this case it is uniform at zero), the region must be charge neutral. This neutral region is referred to as the current-induced base, and has width c09-math-0316 as indicated in Figure 9.9. Since the field is zero in this region, there can be no electron drift, and the collector current is supported entirely by electron diffusion. Since diffusion requires a concentration gradient, and since the region must remain charge neutral throughout, additional holes must be present in a concentration to exactly balance the additional electrons at each point. Electron diffusion in a charge-neutral region is the same process that takes place in the neutral base and, therefore, the current-induced base acts as an extension of the metallurgical base. This effect is referred to as base push-out, and results in a reduction in the current gain in the forward-active region at high current densities.

9.1.8 Common Emitter Current Gain: Temperature Dependence

The common emitter current gain of a narrow-base npn BJT under low-level injection was given in Equation 9.19, where the density of ionized acceptors in the base c09-math-0317 appears in the denominator. As discussed in Appendix A, aluminum acceptors in 4H-SiC have an ionization energy around 200 meV and not all the acceptors are ionized at room temperature. Figure A.1 shows that at a doping density of c09-math-0318 only about 30% of the acceptors in the base are ionized and contribute holes at 23 °C. The low density of holes in the base increases both the emitter injection efficiency (Equation 9.18) and c09-math-0319 (Equation 9.19). However, as temperature goes up, a larger fraction of the acceptors become ionized, and this causes c09-math-0320 to decrease. For a BJT with base doping of c09-math-0321, the common emitter current gain will fall to one-third of its room temperature value at 300 °C. For this reason, it is important to evaluate the performance of the BJT, indeed all SiC power devices, at the highest junction temperature envisioned for the particular application.

The decrease in c09-math-0322 with temperature is actually beneficial in one respect, since it helps prevent thermal runaway. This makes it possible to parallel multiple BJTs in high-current modules.

9.1.9 Common Emitter Current Gain: the Effect of Recombination

The current gain in SiC BJTs is profoundly affected by recombination taking place in the base and emitter regions, at the surfaces of the base and emitter, at the EB and CB junctions, and in the c09-math-0323 implanted regions typically used to facilitate base contacts. We will discuss each of these effects below.

Recombination (and generation) is treated in most standard device textbooks. We are concerned here with recombination through deep levels in the bandgap associated with crystal defects, a process known as Shockley–Read–Hall (SRH) recombination. The net recombination rate per unit volume through single-level SRH centers in a bulk crystal can be written

where c09-math-0325 and c09-math-0326 are hole and electron capture cross sections, c09-math-0327 is the thermal velocity (c09-math-0328 at room temperature), c09-math-0329 is the density of SRH centers per unit volume, c09-math-0330 and c09-math-0331 are the hole and electron minority carrier lifetimes, and c09-math-0332 and c09-math-0333 are given by

9.76 equation

where c09-math-0335 is the energy of the SRH center in the bandgap. For midgap centers, c09-math-0336 and c09-math-0337 are each equal to c09-math-0338.

With appropriate modifications, Equation 9.75 can also be applied to two-dimensional crystal surfaces or interfaces where we have a distribution of states with respect to energy across the bandgap. In these cases, the net recombination rate per unit area can be written

where c09-math-0340 is the density of recombination centers per unit area at the surface or interface, and c09-math-0341 and c09-math-0342 are the surface recombination velocities for holes and electrons, respectively. In both Equations 9.75 and 9.77 the net recombination rate goes to zero in equilibrium c09-math-0343, and becomes negative (net generation) when c09-math-0344. Both surface and bulk recombination rates are proportional to the density of SRH centers c09-math-0345 or c09-math-0346, so minimizing the density of such centers is essential to reducing recombination.

In SiC BJTs, recombination events remove electrons injected from the emitter into the base, preventing them from contributing to collector current. Recombination also removes holes injected from the base into the emitter. This increases the gradient of hole density in the emitter, which increases the base current. Both of these effects reduce c09-math-0347.

The dominant recombination sites in the BJT are illustrated schematically in Figure 9.10. Recombination can occur through defects in the neutral base and neutral emitter, indicated by (a) in the figure. We will discuss this recombination momentarily. Surface recombination can occur at the top surfaces of the base and emitter, and at the sidewalls of the emitter, labeled (b) in the figure. To minimize surface recombination, it is important to employ the best possible surface passivation, typically a thermal or deposited oxide followed by post-oxidation anneal in nitric oxide [1]. In addition, orienting the emitter fingers so their sidewalls lie on the c09-math-0348 plane can reduce emitter sidewall recombination, since this plane has a lower surface recombination velocity [1].

c09f010

Figure 9.10 Illustration of the important recombination sites in a SiC power BJT.

Recombination can also occur in the implanted c09-math-0349 regions under the base contacts, region (c) in the figure, due to crystal damage caused by the implantation [2]. An effective solution is to locate the c09-math-0350 implants several diffusion lengths (or several base widths, whichever is shorter) away from the emitter edge. Recombination also occurs through defects at the pn junctions, shown as (d) in the figure, especially when the epigrowth is interrupted at these interfaces. This recombination can be reduced by growing the entire npn structure using continuous epigrowth [1]. Finally, we must consider recombination at the emitter ohmic contact, region (e) in the figure. This increases the hole current injected from the base and reduces c09-math-0351, but this degradation can be minimized by making the emitter thicker than the hole diffusion length c09-math-0352.

Since surface recombination occurs along or close to the emitter edges, c09-math-0353 can often be increased by using wide emitter fingers, thereby decreasing the perimeter-to-area ratio of the emitter [1]. However, this has to be balanced against the increase in specific on-resistance c09-math-0354. At first glance, increasing the emitter area relative to cell area would seem to make the cell more efficient, but we need to consider the effect of base spreading resistance. The lateral sheet resistivity of the base is given by

This resistance can be significant in SiC BJTs due to the low hole mobility c09-math-0356 and low acceptor ionization percentage c09-math-0357. Base current flowing from the base contact through the lateral resistance of the base produces a lateral voltage drop that reduces c09-math-0358 under the center of the emitter, reducing both electron and hole currents c09-math-0359 and c09-math-0360. This does not affect c09-math-0361, since c09-math-0362 and c09-math-0363 are reduced by the same factor, but it makes the interior of the emitter inactive and increases the on-resistance. In choosing the width of the emitter, the designer must evaluate the trade-off between enhancing c09-math-0364 and degrading c09-math-0365.

Process (a) in Figure 9.10 represents SRH recombination within the neutral base and neutral emitter, and this can obviously limit c09-math-0366. The dominant SRH centers in bulk 4H-SiC are c09-math-0367 centers associated with carbon–silicon divacancies, and c09-math-0368 centers associated with carbon vacancies. These deep levels can be minimized by a 5-h thermal oxidation at 1150 °C prior to the implant activation anneal, followed by another 5-h oxidation after the implant activation anneal [1]. SiC power BJTs receiving such anneals, along with the other measures discussed above, have demonstrated c09-math-0369 above 250 at room temperature [1].

9.1.10 Blocking Voltage

In the previous sections we considered the on-state performance of the BJT. We now turn to the blocking voltage. In the blocking state, the CB junction is reverse biased and supports the entire collector voltage. The maximum voltage that can be applied to the collector is limited either by punch-through of the neutral base or by avalanche breakdown of the CB junction.

Punch-through occurs when the depletion region of the CB junction extends across the neutral base and merges with the depletion region of the EB junction. When this occurs, the potential barrier confining electrons to the emitter is reduced and a large electron current flows from emitter to collector. The current is no longer controlled by the base, since the neutral portion of the base has vanished. As discussed in Section 10.1, punch-through can be prevented by insuring that the doping-thickness product of the base is large enough that the base cannot be completely depleted before the onset of avalanche breakdown at the CB junction. This condition is given by Equation 10.1. Since Equation 10.1 is satisfied in a well-designed BJT, the blocking voltage will normally be limited by avalanche breakdown.

Avalanche breakdown and edge termination techniques are discussed in Chapter 10, but there is one aspect of the BJT that requires special attention, namely the role of the internal current gain of the BJT. In the blocking state the base current is held at zero, which is equivalent to an open-circuit condition at the base terminal. As will be discussed in Section 10.1.1, avalanche breakdown is the result of impact ionization of carriers in the high-field region of the reverse-biased CB junction. Impact ionization generates electron–hole pairs in the depletion region, and the electric field separates the carriers, drawing the electrons into the collector and holes into the base. With c09-math-0370 set to zero, these holes cannot flow out the base terminal and instead must flow into the emitter. This is equivalent to an internal base current that is amplified by the gain of the BJT. In other words, for every hole flowing into the emitter, approximately c09-math-0371 electrons are injected from the emitter into the base. These electrons diffuse across the base and are swept into the CB depletion region where they initiate additional impact ionization. The holes generated by these new ionization events are themselves swept toward the emitter where they cause the injection of even more electrons, and so on. Once initiated, this process increases without bound, resulting in an uncontrolled increase in collector current at a reverse voltage that would be insufficient to cause avalanche breakdown in an isolated CB junction. As discussed in many textbooks, the blocking voltage of the BJT with base open-circuited c09-math-0372 is related to the breakdown voltage of an isolated CB junction c09-math-0373 by the equation

where c09-math-0375 is a constant, typically between 3 and 6. Since the denominator in Equation 9.79 is greater than unity, the open-base blocking voltage of the BJT is less than the breakdown voltage of the CB junction in isolation.

..................Content has been hidden....................

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