Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-27T00:29:53.107Z Has data issue: false hasContentIssue false

Modeling of a GaN Based Static Induction Transistor

Published online by Cambridge University Press:  13 June 2014

Gabriela E. Bunea
Affiliation:
Dept. of Physics, Boston University, Boston, MA, 02215, gbunea@bu.edu
S.T. Dunham
Affiliation:
Dept. of Electrical and Computer Engineering, Boston University, Boston, MA, 02215
T.D. Moustakas
Affiliation:
Dept. of Electrical and Computer Engineering, Boston University, Boston, MA, 02215

Abstract

Static induction transistors (SITs) are short channel FET structures which are suitable for high power, high frequency and high temperature applications. GaN has particularly favorable properties for SIT operation. However, such a device has not yet been fabricated. In this paper we report simulation studies on GaN static induction transistors over a range of device structures and operating conditions. The transistor was modeled with coupled drift-diffusion and heat-flow equations. We found that the performance of the device depends sensitively on the thermal boundary conditions, as self-heating effects limit the maximum voltage swing.

Type
Research Article
Copyright
Copyright © 1999 Materials Research Society

Introduction

GaN is a wide-bandgap semiconductor (Eg=3.4 eV), and therefore has a high breakdown field [Reference Shur and Khan1] and low thermal generation rate. These properties combined with good thermal conductivity and stability make GaN an attractive material for high power/ high temperature and radiation harsh environment electronic devices. Monte Carlo simulations predict a peak electron velocity of 3.2×107 cm/s and a saturation electron velocity of 2.5×107 cm/s [Reference Khan, Chen, Shur, Dermott, Higgins, Burm, Schaff and Eastman2]. This makes possible high frequency operation of GaN devices.

SIT’s are short channel FET structures in which the current, flowing vertically between source and drain, is controlled by the height of an electrostatically induced potential barrier under the source [Reference Siergiej and Clarke3]. A cross-sectional diagram of the SIT is shown in Figure 1.

Figure 1. Cross-section of static induction transistor (SIT) structure considered in this work and the unit cell simulated with critical dimensions labeled.

Electrons are emitted from the source, which is at ground potential, and are accelerated to the drain, which is biased at positive potential, where they are collected [Reference Moore and Trew4]. A very thin heavily doped layer is deposited next to the drain and source contacts in order to form ohmic contacts. A grid structure is located in the space between the source and drain electrodes so the charged carriers can be externally modulated. The RF gain of the device is determined by the efficiency with which the modulation is affected. The grid structure is generally fabricated using pn or Schottky junctions.

A range of field effect transistors including MESFET, MISFET, inverted channel AlGaN/GaN and MODFET have been developed with potential applications for high power/ high temperature electronics [Reference Khan, Chen, Shur, Dermott, Higgins, Burm, Schaff and Eastman2,Reference Binari5-Reference Burm, Schaff, Martin, Eastman, Amano and Akasaki8]. To our knowledge, the highest cut-off frequency reported for GaN-based FET’s is 52 GHz [Reference Wu, Keller, Keller, Fini, Pusl, Le, Nguyen, Nguyen, Widman, Denbaars, Keller and Mishra9], and the maximum frequency of oscillations is over 97 GHz [Reference Khan, Chen, Shur, Dermott, Higgins, Burm, Schaff and Eastman10]. Significant results regarding the power output of GaN-based FET’s have been reported by several groups [Reference Wu, Keller, Keller, Fini, Pusl, Le, Nguyen, Nguyen, Widman, Denbaars, Keller and Mishra9, Reference Wu, Keller, Keller and Nguyen11]. Wu reported an output power of 3 W/mm at 18 GHz, with a power added efficiency (PAE) of 19% for a 0.25 μm gate AlGaN/GaN MODFET [Reference Wu, Keller, Keller, Fini, Pusl, Le, Nguyen, Nguyen, Widman, Denbaars, Keller and Mishra9]. In comparison, our SIT simulation results show a cut-off frequency fT of 24.8 GHz, a maximum frequency of oscillations fmax of 75.2 GHz. Operated under class B, the output power decreases from 10.75 W/mm to 1.95 W/mm as the operating frequency changes from 2 GHz to 40 GHz. Correspondingly, the PAE changes from 73.8% to 9.1 %.

Theory

Basic simulation equations and physical models

We employed a commercially available 2D device simulator (ATLAS [12]) which was modified appropriately for GaN, based on experimental observations and theoretical calculations. The transistor was modeled with coupled drift-diffusion and heat-flow equations. The effect of lattice temperature on the performance of the device was taken into account by including the thermoelectric factor in current density equations (1,2) and by adding the heat-flow equation (3).

(1)
(2)
(3)

where T is the lattice temperature, and Pn and Pp are thermoelectric power coefficients for electrons and holes, respectively, C is the heat capacitance per volume (1.97 J/K cm3 [Reference Anderson13]), κ is the thermal conductivity (1.3 W/cm K [Reference Sichel, Pankove and Phys14]) and H is the heat generation based on Joule effect.

The models used in the simulation are based on those from Si and GaAs, but have been modified to fit the available data for GaN. The temperature dependence of bandgap energy:

(4)

with Eg(0)=3.5 eV, α=9.39×10−4 eV/K, and β=772 K, based on optical absorption measurements [Reference Dmitriev and Oruzheinikov15]. The electron (hole) low-field mobility as a function of the impurity concentration (N) and temperature T is given by [12]:

(5)

where μ1, μ2, β, γ,δ and Ncrit were determined by fitting the values available from the literature for both electrons and holes [Reference Khan, Chen, Shur, Dermott, Higgins, Burm, Schaff and Eastman2,Reference Albrech16-Reference Götz, Johnson, Chen, Liu, Kuo and Imler19]. For electrons: μ 1 =15 cm2/Vs, μ 2 =1800 cm2/Vs, β= -3.04, γ= -2.55, δ = 0.66, Ncrit = 8×1016 cm−3. For holes: μ1 =0.14 cm2/Vs, μ2 =880 cm2/Vs, β= -1.5, γ=0, δ = 0.67, Ncrit = 5.5×1014 cm−3. A comparison between experimental and predicted values (from Monte Carlo simulation) of low field electron mobility versus doping level and the model used in our simulations is presented in Figure 2a. Figure 2b presents a similar comparison for the electron low field mobility versus temperature [Reference Götz, Johnson, Chen, Liu, Kuo and Imler19]. In our simulations, we employ a channel doping of 5×1016 cm−3. The low field electron mobility reported for such a doping level is 900 cm2/Vs [Reference Binari5,Reference Nakamura, Mukai and Senoh20]. However this mobility was measured laterally. There is evidence that the lateral mobility is reduced due to the scattering by charged dislocations, while the vertical mobility is significantly higher because the electrons are repelled from the dislocation lines by band bending due to the negative charge on the dislocations [Reference Weimann, Eastman, Doppalapudi, Ng and Moustakas21]. In our device, which is a vertical one, we employ a room temperature electron mobility of 1050 cm2/Vs, which we believe is a reasonable value.

Figure 2. Electron mobility (a) versus doping level at T=300K [Reference Khan, Chen, Shur, Dermott, Higgins, Burm, Schaff and Eastman2, Reference Albrech16-Reference Götz, Johnson, Chen, Liu, Kuo and Imler19], and (b) versus temperature, for a doping level of 3×1017 cm−3 [Reference Götz, Johnson, Chen, Liu, Kuo and Imler19].

For high fields, we use a simple model for mobility versus electric field, which ignores the overshoot effect. This slightly underestimates the current, but the effect is minor for high power devices because they operate at high fields. In high electric fields, the saturation velocity is weakly dependent on temperature. The saturation velocity is modeled as a function of temperature by an empirical relation obtained by fitting the results from Monte Carlo simulation [Reference Bhapkar and Shur18]:

(6)

We assumed that optical recombination is given by:

(7)

with Copt =3×10−11cm3/s, from absorption experimental data and calculated electron energy band dispersion [Reference Dmitriev and Oruzheinikov15].

The generation rate of electron-hole pairs due to impact ionization is modeled according to Selberherr [Reference Selberherr22]:

(8)

To our knowledge, there are no experimental measurements of impact ionization rates on GaN. However, calculations of impact ionization rates using ensemble Monte Carlo simulation including the full details of all the relevant valence bands, based on pseudopotential approach, have been published recently [Reference Oguzman, Bellotti, Brennan, Kolnik, Wang and Ruden23]. We assume that the impact ionization rates (αn and αp ) are dependent on the electric field and temperature for both electrons and holes according to the formula [12, Reference Selberherr22]:

(9)

The critical electric field also depends on T:

(10)

By fitting the results from [Reference Oguzman, Bellotti, Brennan, Kolnik, Wang and Ruden23], the parameters were found to be: for electrons (α = 4.55 × 106 cm−1, E0 crit =1.19×107V/cm, β=1), for holes (α = 1.48 × 106 cm−1, E0 crit =8.95×106 V/cm, β=1). Since there are no available results for T dependence of . we use the values for silicon for both electrons and holes (A=0.588, B=0.248, M=1 [12]).

Device optimization

In order to optimize the SIT structure for operation at high power, high temperature and high frequency, DC, small signal and large signal analysis have been performed. Due to symmetry, we need to simulate only one unit cell of the transistor (Figure 1). In order to maximize the breakdown voltage as well as carrier mobility we want low doping in the channel, and 5×1016 cm−3 was chosen as a value that is achievable with current technology. As the distance between source and gate (dSG ) decreases, the voltage gain μ as well as transconductance, gm increases. The value of dSG is limited by the need to avoid the source to gate punch-through, when the depletion region from gate extends to that from the source. A large distance between gate and drain (dGD ) is desirable in order to have a large breakdown voltage. However as dGD increases the series resistance also increases and this limits the frequency and current response of the device. The half-width of the source finger (WS ) determines the blocking voltage of the transistor, so it controls the voltage swing in power measurements. By balancing the above considerations we obtained the following values: the channel is 3 μm thick with a doping concentration of 5×1016 cm−3; the n+ layers are 0.2 μm thick and have a doping concentration of 1×1019 cm−3. The top of the device is modified in a comb configuration having the following dimensions: source half-width WS =0.5 μm, gate half-width WG =1.5 μm, gate height hG =0.2 μm, dSG =0.6 μm, dGD =2.4 μm.

Results

Our simulations indicate that the performance of the device is very sensitive to the thermal boundary conditions, as self-heating effects limit the voltage swing. In order to increase the power output we have assumed that the device is build on a SiC substrate. According to Binary [Reference Binari5] the SiC substrate allows about a 4x increase in power density due to the higher thermal conductivity of SiC compared to GaN or sapphire. For the same purpose, a layer of diamond paste is assumed to be present on the top of the device. We have calculated the equivalent thermal impedance of the SiC substrate and diamond layer, assuming that the length and height of the SIT is much smaller than the thickness and length of the SiC and diamond layers [12,Reference Hahne and Grigull24]. For one finger of the device, the calculations lead to a thermal impedance of 3300 W/cm2 K for drain contact, 12000W/cm2 K for source contact and 4050 W/cm2 K for gate contact. Simulated drain I-V characteristics with the gate voltages varying from 0V to -12V are shown in Figure 3a. We notice that the breakdown voltage varies with the gate bias from about 50V at 0V gate bias to 320V at −12V gate bias. The simulations indicate that the breakdown is due to self-heating effects. Our results show that the performance of the device is not significantly affected by thermal generation as long as the maximum temperature in the device does not exceed 700K. The gains obtained by small signal ac simulations for a drain bias of Vd=100V and a gate bias of Vg=-6V are plotted as a function of frequency in Figure 3b. The cut off frequency is fT=24.8 GHz and the maximum frequency of oscillations is fmax=75.2 GHz.

Figure 3. Transistor characteristics: (a) drain I-V characteristics, and (b) gain versus frequency. The cut-off frequency is fT=24.8 GHz and maximum frequency of oscillations fmax=76.5 GHz.

Large signal analysis was performed under class B, in order to obtain the output power and power added efficiency (PAE). In Figure 4 we present the output power and PAE as functions of frequency, with a DC drain bias of VDD=180V, load resistance of 7.2×105 ohm and a gate bias voltage swing between 0 and −12 V. Under these conditions, the output power varies from 10.75 W/mm to 1.95 W/mm and PAE varies from 73.8 % to 9.1 % as we increase the operation frequency from 2 GHz to 40 GHz. The maximum theoretical PAE for operation under class B is 78.5 % [Reference Moore and Trew4]. Note that operation under class B requires two transistors and gate periphery of both transistors is included in output power calculations.

Figure 4. Output power and power added efficiency (PAE) versus frequency.

Conclusions

In conclusion, we modeled a static induction transistor based on GaN films. Our results show that with a SiC substrate and top side diamond paste as thermal sinks, the output power can be as high as 10.75 W/mm at 2 GHz operating frequency, with a PAE of 73.8 %. The cut-off frequency was found to be 24.8 GHz and the maximum frequency of oscillations was 75.2 GHz. These results demonstrate the excellent potential of GaN based static induction transistors for high power, high temperature and high frequency operation.

This work was supported in part by ONR through a subcontract by Raytheon.

References

Shur, M.S., Khan, M.A., Mat. Res. Bull. 22 (2), 44 (1997).CrossRefGoogle Scholar
Khan, M.A., Chen, Q., Shur, M., Dermott, B., Higgins, J., Burm, J., Schaff, W., Eastman, L., Solid State Electron. 41, 1555 (1997).CrossRefGoogle Scholar
Siergiej, R.R., Clarke, R.C., IEDM-95, 353 (1995).CrossRefGoogle Scholar
Moore, K., Trew, R.J., MRS Bulletin, March, 50 (1997).CrossRefGoogle Scholar
Binari, S., The 2nd Int. Conf. On III-V Nitrides, Tokushima, Japan (1997).Google Scholar
Trew, R.J., Shin, M., Gatto, W., Solid State Electron. 41, 1561 (1997).CrossRefGoogle Scholar
Wu, Y.F., Keller, B., Keller, S., Kapolnek, D., Kozodoy, P., Denbaars, S., Mishra, U., Solid State Electron. 41, 1569 (1997).CrossRefGoogle Scholar
Burm, J., Schaff, W., Martin, G., Eastman, L., Amano, H., Akasaki, I., Solid State Electron. 41, 247 (1997).CrossRefGoogle Scholar
Wu, Y.F., Keller, B., Keller, S., Fini, P., Pusl, J., Le, M., Nguyen, N., Nguyen, C., Widman, D., Denbaars, S. Keller, , , S., Mishra, U., Electron. Lett. 33, 1742 (1997).CrossRefGoogle Scholar
Khan, M.A., Chen, Q., Shur, M., Dermott, B., Higgins, J., Burm, J., Schaff, W., Eastman, L., IEEE Electron Device Lett. 17, 584 (1996).CrossRefGoogle Scholar
Wu, Y.F., Keller, B., Keller, S., Nguyen, N., IEEE Electron Device Lett. 18, 438 (1997).CrossRefGoogle Scholar
Atlas User’s Manual (Device simulation software), Silvaco Inc., Version 1.5.0. (1997).Google Scholar
Anderson, H., Physics Vade Mecum (AIP, 1981).Google Scholar
Sichel, E.K., Pankove, J.I., Phys, J.. Chem. Solids 38, 330 (1978).CrossRefGoogle Scholar
Dmitriev, A.V., Oruzheinikov, A.L., MRS Internet J. Nitride Semic. Res. 1, 46 (1996).CrossRefGoogle Scholar
Albrech, J.D., MRS Nitride Symp. Proc., 423 (1996).Google Scholar
Shur, M., J. Electron. Mat. 25, 777 (1996).CrossRefGoogle Scholar
Bhapkar, U., Shur, M.S., J. Appl. Phys. 82 (4), 1649 (1997).CrossRefGoogle Scholar
Götz, W., Johnson, N.M., Chen, C., Liu, H., Kuo, C., Imler, W., Appl. Phys. Lett. 68 (22), 3144 (1996).CrossRefGoogle Scholar
Nakamura, S., Mukai, T., Senoh, M., Jpn. J. of Appl. Phys. 31, 2883 (1997).CrossRefGoogle Scholar
Weimann, N., Eastman, L., Doppalapudi, D., Ng, H., Moustakas, T.D., J. Appl. Phys. 83, 3656 (1998).CrossRefGoogle Scholar
Selberherr, S., Analysis and simulation of semiconductor devices (Wien-New York, 1984).CrossRefGoogle Scholar
Oguzman, I.H., Bellotti, E., Brennan, K., Kolnik, J., Wang, R., Ruden, P., J. Appl. Phys. 81 (12), 7827 (1997).CrossRefGoogle Scholar
Hahne, E., Grigull, U., Int. J. Heat Mass Transfer 18, 751 (1975).CrossRefGoogle Scholar
Figure 0

Figure 1. Cross-section of static induction transistor (SIT) structure considered in this work and the unit cell simulated with critical dimensions labeled.

Figure 1

Figure 2. Electron mobility (a) versus doping level at T=300K [2, 16-19], and (b) versus temperature, for a doping level of 3×1017 cm−3 [19].

Figure 2

Figure 3. Transistor characteristics: (a) drain I-V characteristics, and (b) gain versus frequency. The cut-off frequency is fT=24.8 GHz and maximum frequency of oscillations fmax=76.5 GHz.

Figure 3

Figure 4. Output power and power added efficiency (PAE) versus frequency.