Modeling Effects Of Electron-Velocity Overshoot In A MOSFET

advertisement
IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 44, NO. 5, MAY 1997
841
Modeling Effects of Electron-Velocity
Overshoot in a MOSFET
J. B. Roldán, F. Gámiz, Member, IEEE, J. A. López-Villanueva, Member, IEEE, and J. E. Carceller, Member, IEEE
Abstract— A simple analytical expression to account for
electron-velocity overshoot effects on the performance of very
short-channel MOSFET’s has been obtained. This new model
can be easily included in circuit simulators of systems with a
huge number of components. The influence of temperature and
low-field mobility on the increase of MOSFET transconductance
produced by electron-velocity overshoot as channel lengths
are reduced can be easily taken into account in our model.
The accuracy of this model has been verified by reproducing
experimental and simulated data reported by other authors.
I. INTRODUCTION
N
ONLOCAL effects are becoming more and more
prominent as MOSFET dimensions shrink to the deepsubmicrometer regime. Velocity overshoot is one of the
most important effects from the practical point of view as
it is directly related with the increase of current drive and
transconductance experimentally observed in short-channel
MOSFET’s [1]–[4]. Some authors [1]–[4] have shown
that experimental measurements of submicron MOSFET
transconductances are higher than the theoretical maximum
transconductance that can be reached in the case where
electrons drift in equilibrium with the lattice, with their
electron velocity being limited by the saturation velocity
[1], [5]. This result has been shown for channel lengths
under 0.15 m, which means that the electron velocity along
the channel is higher than the saturation one; therefore, if
velocity overshoot can be controlled, the performance of very
short-channel MOSFET’s can be improved with respect to
the performance of long-channel transistors. This effect is
also reasonably well understood from the theoretical point
of view [6], [7]. It has been shown that an electric field
step causes the electron velocity to overshoot the value that
corresponds to the higher field for a period shorter than the
energy relaxation time [6], [7] (the time needed by the electron
to once again reach equilibrium with the lattice [8]). Therefore,
as the longitudinal electric field increases, the electron gas
starts to be in disequilibrium with the lattice [9]. There is an
insufficient number of phonon-scattering events experienced
by the electron during its flight, with the result that electrons
can be accelerated to velocities higher than the saturation
velocity, thus approaching ballistic transport conditions.
This effect is due to the nonequivalence of momentum and
energy-relaxation times and can be observed for a period
shorter than the energy relaxation time. Hence, overshoot is
Manuscript received October 1, 1996. The review of this paper was arranged
by Editor D. P. Verret. This work has been carried out within the framework
of research project TIC 95-0511, supported by the Spanish Government
(CICYT).
The authors are with the Departamento de Electrónica y Tecnologı́a de
Computadores. Universidad de Granada, 18071 Granada, Spain.
Publisher Item Identifier S 0018-9383(97)03009-8.
a clear nonequilibrium effect and cannot be predicted with
simple drift-diffusion simulators, it being instead necessary
to use more sophisticated simulators capable of dealing with
nonequilibrium transport, such as the hydrodynamic [7] or
Monte Carlo (MC) [9]–[11] simulators. In fact, the high value
of the transconductance for transistors with channel lengths
under 0.15 m in an MC simulation [9] has been explained as
being a consequence of nonequilibrium transport in the source
edge. The same results were achieved by a drift-diffusion
simulation augmented to account for velocity overshoot by
means of a mobility model that depends on the electron
temperature obtained under energy conservation conditions
[12].
The need for more sophisticated device simulations is
becoming an easily solvable problem, as computer speed and
storage capacity are rapidly increasing. Nevertheless, for this
same reason it is easy to foresee that simple drift-diffusion
simulators will likely be included in circuit computer-aideddesign tools instead of the simpler analytical models in the near
future in order to increase simulation accuracy. In addition,
the simple analytical model can still be used for simulation
of systems including a huge number of components. In short,
there is still a need for accurate models at different levels,
but as the trend of MOSFET technology is toward sub-0.1 m
[1], [13], it is desirable to be able to predict velocity-overshoot
effects at these different modeling levels.
The aim of this study is to model velocity-overshoot effects
on MOSFET transconductance. MOSFET transconductance,
, is known to be the most important figure of merit in
dealing with the large-signal switching performance of logic
devices, as the time constant for a small MOSFET to charge a
load is proportional to
, where is the node capacitance.
That is why accurate modeling of this parameter is essential
in circuit simulators of state-of-the-art MOSFET’s [14].
We have used an augmented drift-diffusion velocity model
reported previously by several authors [15]–[18] to obtain
a simple analytical expression to model the MOSFET
transconductance increase that has been experimentally
observed [1]–[4] as MOSFET effective channel lengths shrink.
The results will be compared to the experimental and the MC
simulated transconductances obtained by G. A. Sai-Halasz et
al. [1] and M. R. Pinto et al. [19], respectively. Low-field
mobility and temperature effects can be easily taken into
account, as will be shown.
II. TRANSCONDUCTANCE CALCULATION
We suppose that the source is located at
and the
potential at that point is zero; the drain is assumed to be at
(where is the effective channel length, henceforth we
0018–9383/97$10.00  1997 IEEE
842
IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 44, NO. 5, MAY 1997
will mean effective channel length wherever we use channel
length) at a potential
The drain current can be calculated
as a product of the mobile channel charge, the channel width,
and the electron velocity along the
axis, from source
toward the drain. Rearranging this relationship to obtain a term
multiplied by
and a term multiplied by
, and integrating
these terms along the channel, from 0 to and from 0 to
,
respectively, yields an expression for the drain current.
The drift velocity in inhomogeneous electric field can be
expressed approximately as [16]
We integrated the above expression making use of the meanvalue theorem for the second term in brackets and a simple
variable change for the rest. The result is given by
(1)
where the mean value is shown in angles for the expression
above, and the function
is the integral of the
inversion charge in the channel
where
is the drift velocity in a homogeneous field. The “ ”
parameter was supposed by Thornber [16] to be dependent
on the longitudinal-electric field. Artaki also studied this
dependence in detail [17]. In both cases, this dependence
was shown to be important at low-longitudinal-electric fields.
However, the higher values of the longitudinal-electric-field
gradient are located near the drain edge in short-channel
MOSFET’s, where high longitudinal fields can also be found
under normal operation conditions. The dependence of the
parameter on the longitudinal field is not very important for
these high-field gradients. In order to check this fact, we have
performed many MC deep-submicron MOSFET simulations
including inversion layer quantization by changing the channel
length and bias. The results showed that the parameter can
be regarded as a constant for the usual longitudinal field range
found in short-channel MOSFET’s. The reduced complexity of
(1) when the term connected to the longitudinal-field gradient
is taken as a constant allows us to obtain a closed analytical
expression for the transconductance.
A useful expression (2) has been given for the drift velocity
in homogeneous field
[20]
(2)
At this stage some simplifying assumptions are made: we
take an average value for the low longitudinal-electric-field
mobility corresponding to an average value of the transverseelectric field along channel , which is not a rough approximation for low- and medium-drain voltages. We also take
, which simplifies the algebra below. This approximation
is along the lines of a series expansion of the denominator in
(2) proposed by N. D. Arora [21], and has also been reported
by other authors [22], [23].
The electron-velocity expression under the reported assumptions has been put into the drain-current equation [5].
Integrating from source to drain along the channel we get
(4)
(5)
On one hand, the expression in angles is very high near the
drain compared to the value of this expression in other zones
of the channel. On the other hand, the value of the inversion
charge which multiplies this value in the integral is higher near
the source than in the rest of the zones closer to the drain, so
there is no zone of the channel where the integrand is clearly
dominant. Therefore, we will use a semiempirically modified
mean value along the channel for the expression in angles. To
do so, we approximate the first and second derivative of the
electrostatic potential as follows:
(6)
where is a constant that could be dependent on the operation
voltages and the technological features of the MOSFET’s such
as the low-field mobility. An expression similar to the first
part of (6) can be found in [24]. The approximations shown
in (6) and some others explained above have to be considered
within the first-order approximation empirical approach we are
dealing with in order to obtain a closed analitycal expression
to model electron-velocity-overshoot effects in short-channel
MOSFET’s. As can be seen, both the
and
parameters
contribute to increase the second term in the same way, so the
dependencies of both parameters can be merged in a single
one, , which will be used henceforth. The final result we
obtained for the drain current after the previous considerations
was given by
(7)
Making use of (7), the MOSFET transconductance can be
calculated as
(8)
(3)
(Henceforth transconductances adjusted for width normalization will be used.)
In accordance with the above calculation, it is obvious
that the quotient of the transconductance, accounting and
ROLDAN et al.: MODELING EFFECTS OF ELECTRON-VELOCITY OVERSHOOT
843
Fig. 1. Room-temperature ratio between experimental and simulated
intrinsic transconductances versus channel length for a sample of MOSFET’s
reported by G. A. Sai-Halasz et al. [1] is shown in squares for a
VT
0:6 V, VDS = 0:8 V. The same ratio
bias of VGS
calculated analytically by (9) with the following values: VDS = 0:8 V,
2
m = 390 cm /Vs; vsat = 8 106 cm/s; and a = 25 1005 cm3 =Vs
is shown in solid line.
Fig. 2. Low-temperature (T = 77 K) ratio between experimental and
simulated intrinsic transconductances versus channel length for a sample of
MOSFET’s reported by G. A. Sai-Halasz et al. [1] is shown in squares for
VT = 0:6 V, VDS = 0:8 V. The same ratio
a polarization of VGS
calculated analytically by (9) with the following values: VDS = 0:8 V,
2
= 720 cm =Vs; vsat = 1 107 cm/s, and a = 40 1005 cm3 =Vs
is shown in solid line.
0
=
2
2
not accounting for the effects of velocity overshoot, can be
calculated as
(9)
III. RESULTS
AND
DISCUSSION
In order to check this expression, we have calculated
the ratio between experimental intrinsic transconductances
measured for a sample of MOSFET’s fabricated by G. A. SaiHalasz et al. [1] and the same transconductances calculated by
these authors using a classical drift-diffusion simulator with
V,
V,
cm Vs
cm/s at
K. The ratio provided in reference [1]
has been plotted in Fig. 1 in squares. The same ratio, calculated
using (9) for the same parameters, is shown in solid line for
cm Vs and the constant values given by SaiHalasz et al. An excellent agreement is reached throughout the
range of effective channel lengths.
Value
is coherent with the significant increase of the
saturation velocity reported in [1] that the authors needed to
use in the drift-diffusion simulator to reproduce the experimental transconductances; a significant increase of the saturation
velocity for both room temperature and low temperature was
needed, mostly for channels shorter than 0.1 mm. Examining
(1), the first term corresponding to the homogeneous transport
regime is found to saturate at high longitudinal-electric fields,
so a value of 25–40% of the saturation velocity is expected
for the second term of the equation in order to obtain the same
velocities reported by the authors at the lowest channel lengths.
To get an average increase of 25–40% in the electron velocity
along the channel, particulary for the saturation velocity, a
value of
times
has to be introduced in
(1) in place of
, and therefore the value found for
parameter
is suitable.
The same verification procedure was repeated for
K in Fig. 2. A perfect fit, as before, was obtained using the
parameters reported in [1] for the simulation of the MOSFET’s
at low temperature, which were
V,
V,
cm Vs,
cm/s
0
2
2
cm Vs An increase of the
value is coherent as we
reduce temperature since as temperature goes down electronphonon interaction is lower than at room temperature [25].
As can be observed in Fig. 2, the ratio of experimental and
drift-diffusion simulated transconductances is higher at low
temperatures for all channel lengths for the same devices,
which supports the explanation given above. The same reasoning used to determine the appropriate value for parameter ,
taking into account similar values for the longitudinal-electricfield derivative (the same bias is used at both temperatures),
can be applied in this case. The increase of saturation velocities
used in the drift-diffusion simulator to fit the experimental
transconductances is higher than at room temperature, which
is why a higher value of
is expected, again, to account for
this additional increase.
We have also tried to reproduce the experimental intrinsic
transconductances by applying (8). To do so, we fitted the
measurements by G. A. Sai-Halasz et al. using (8) and an
estimated value for the partial derivative of
To
get this estimated value we took as a starting point a general
expression for the absolute value of the inversion charge per
unit area in MOSFET’s [5]
(10)
where
is the oxide capacity per unit area,
is the flatband voltage,
is the surface potential and
the depletion
charge per unit area. The voltage between substrate and source
was zero. Therefore, we can substitute
by
In
strong inversion, the surface potential can be approximated
by
along the channel, where
equals
is the difference between the Fermi level and
the intrinsic Fermi level in the bulk in eV, and
being Boltzmann’s constant). If we substitute this value of the
surface potential in strong inversion (which is the case here, in
(10) if
is represented only by its value at the source, if its
variation along the channel is neglected, and if the definition
of threshold voltage is used [5], we obtain
(11)
844
IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 44, NO. 5, MAY 1997
Fig. 3. Experimental intrinsic transconductances versus channel length for a
sample of MOSFET’s reported by G. A. Sai-Halasz et al. [1] are shown
VT
0:6 V, VDS = 0:8
in squares for a polarization of VGS
V at T = 300 K and T = 77 K. Intrinsic transconductances calculated making use of (8) with the following values: VDS = 0:8 V,
= 390 cm2 =Vs; vsat = 8 106 cm/s; a = 25 1005 cm3 =Vs;
@F (VGS ; VDS )=@VGS = 5:2 1007 C/cm2 at T = 300 K (dashed line)
and = 720 cm2 =Vs; vsat = 1 107 cm/s; a = 40 1005 cm3 =Vs;
@F (VGS ; VDS )=@VGS = 5:5 1007 C/cm2 at T = 77 K (solid line).
Fig. 4. Simulated intrinsic transconductances versus channel length for
a MOSFET for three different constant channel mobilities reported by
M. R. Pinto et al. [19] are shown in squares (e = 125 cm2 =Vs);
circles (e = 250 cm2 =Vs); and triangles (e = 500 cm2 =Vs)
for a polarization of VGS = VDS = 1:5 V. Intrinsic transconductances calculated making use of (8) with the following values:
vsat = 107 cm/s, @F (VGS ; VDS )=@VGS = 10 1007 C/cm2 and
a = 15; 30; 38 1005 cm3 =Vs for e = 125; 250; 500 cm2 =Vs,
respectively, are shown in solid line.
So, regarding the definition of function
estimated value of the partial derivative of
be easily obtained
the theoretical maximum transconductance
can be
affected by the separation of the inversion charge centroid from
the oxide surface (quantum effects), mostly in short-channel
devices such as the ones fabricated by Ono et al. [13] where
the oxide thickness was 30 Å and the distance of the inversion
charge centroid from the Si/SiO surface can be more than
25% of the oxide thickness. That might be the reason for the
absence of velocity overshoot reported in devices as short as
40 nm [13]. On the other hand, the value of the saturation
velocity of the carriers in the channel is not well known since
reported experimental and simulated values are very different
[11], [28]–[31].
The above discussion reveals it to be desirable to validate
the model by comparison with the results of accurate simulations such as those reported by Pinto et al. [19]. The intrinsic
transconductance versus channel length curves given by Pinto
et al. were calculated by using accurate simulators including
electron-velocity overshoot effects. A common NMOSFET
device was used and a different value for the low-field mobility
was selected for each transconductance curve (Fig. 4). We
have reproduced the curves of Pinto et al. (symbols) by
employing (8) (solid line) and the following parameters:
V,
cm/s (given in [19]) and
C/cm (the same for
all the curves, due to the use of the same device in all
cases). Pinto et al. used different values of the effective
mobility in the channel for their simulations, each different
value corresponding to different technological features in
MOSFET’s. Different
values for each assumed mobility
value have to be employed. The values employed were
cm Vs for
cm Vs,
respectively. It is clear that technological variations play an
important role on the impact of electron-velocity overshoot
effects. As can be seen, these variations can be easily included
in our model.
The expressions deduced for the transconductance (8), and
particularly the term used in (1) to complete Expression
0
=
2
2
2
2
2
2
, (5), an
can
(12)
The value given by G.A. Sai-Halasz et al. for the oxide
width was 4.5 nm, so for the biases reported above
V,
V the estimated value of the partial
derivative of
is
C/cm The values
used to fit the experimental intrinsic transconductances for the
partial derivative for
were
C/cm at
K and
C/cm at 77 K. A perfect fit can
be observed in both cases (Fig. 3) .
The value obtained here is a useful estimation of the partial
derivative of
, but some dependences have been
neglected as the reduction of the channel-inversion charge in
the pinch-off region in the saturation operation region regime,
and other dependences related to application of the meanvalue theorem. These relative errors are comparable to those
reported by G. A. Sai-Halasz et al. in determining channel
lengths due to the difficulty they had in accurately obtaining
them and the series resistance in LDD-like devices with strong
two-dimensional effects [26].
So far, the model has been validated by comparison with
experimental data reported by Sai-Halasz et al. [1]. However,
it is difficult to ascertain the presence of velocity overshoot
from experiments on very short MOSFET’s. In fact, very
critical measurements of gate capacitance and parasitic resistances are required. Therefore, the quantitative estimation of
velocity overshoot from experimental data is likely to present
uncertainties. The intrinsic transconductances are calculated
from the experimental transconductances by using the source
and drain series resistances [27], the uncertainties involved
in the series resistance and effective channel length extraction procedure explained in [26] can significantly reduce the
real transconductance value. In addition, the calculation of
2
2
ROLDAN et al.: MODELING EFFECTS OF ELECTRON-VELOCITY OVERSHOOT
845
(a)
(b)
Fig. 5. Terms A and B described in (13) are shown versus channel length at (a)
in Fig. 3 at each temperature. Term A (white squares), Term B (full squares).
2 accounting for velocity-overshoot effects, can be seen as
a functional approach to the behavior of deep-submicron
MOSFET’s as dimensions shrink to channel lengths under
0.1 m To attempt to see more clearly what the contribution
of the new term to the previous one is as the MOSFET
channel length is reduced, we have plotted them versus channel
length, as they appear in (8), adjusting for channel width
and
normalization. The results for (a)
K and (b)
K are shown in Fig. 5 for the
constant values used in Fig. 3, where Terms A and B are
Term A
Term B
(13)
As can be observed, Term B is not worth taking into account
for channel lengths over 0.2 m at both temperatures. Velocity
overshoot, as expected, does not affect the transconductance
for these lengths. The increase in the saturation-velocity value
needed to simulate the experimental transconductance using
the drift-diffusion simulator reported by G. A. Sai-Halasz et al.
was less than 5% for these lengths; however, as channel length
is reduced, Term B rises, which shows the effects of velocity
overshoot in MOSFET transconductance. Higher modifications
to the saturation velocity were needed to reproduce experimental transconductances using the drift-diffusion simulator
for shorter channel lengths [1]. An obvious greater increase is
expected for saturation velocity needed at
K due to
the higher
parameter (higher low-field mobility) we have
used at this temperature, also reported by G. A. Sai-Halasz
et al. A noticeable increase in MOSFET transconductance has
also been found by other authors in the channel length interval
0.1–0.15 m [9], [12], which is the range where Term B starts
to be important compared to Term A.
It has been shown that the influence of the parameters that
contribute to enhancing velocity-overshoot effects, such as the
longitudinal-electric-field gradient in the MOSFET channel,
temperature, low-field mobility, etc, can be taken into account
by means of (8). Dependences of the
parameter on the
temperature and the low-field mobility are clear; however,
the lack of both experimental and simulated data for a single
T
= 300 K and (b) T
= 77 K for the same constant values used
technology at different biases do not allow bias dependence
on the
parameter to be demonstrated.
IV. CONCLUSIONS
A simple analytical model to account for electron-velocityovershoot effects on submicron MOSFET transconductance
has been provided. It has been shown that as channel dimensions shrink to under 0.15 m, a term inversely proportional
to the square of the MOSFET channel length accurately
describes the transconductance increase produced by electronvelocity overshoot. This model can be easily included in circuit
simulators of systems with a huge number of components. Experimental and simulated data were used to check the accuracy
of our model and to account for the temperature and low-field
mobility dependence of the transconductance increase due to
velocity-overshoot effects. It has been demonstrated that the
lower the temperature and the higher the low-field mobility,
the higher the correction needed to reproduce experimental
and simulated results.
REFERENCES
[1] G. A. Sai-Halasz, M. R. Wordeman, D. P. Kern, S. Rishton, and
E. Ganin, “High transconductance and velocity overshoot in NMOS
devices at the 0.1-m gate-length level,” IEEE Electron Device Lett.,
vol. 9, p. 464, 1988.
[2] S. Y. Chou, D. A. Antoniadis, and H. I. Smith, “Observation of electron
velocity overshoot in sub-100-nm-channel MOSFET’s in silicon,” IEEE
Electron Device Lett., vol. EDL-6, p. 665, 1985.
[3] G. G. Shahidi, D. A. Antoniadis, and H. I. Smith, “Electron velocity
overshoot at room and liquid-nitrogen temperatures in silicon inversion
layers,” IEEE Electron Device Lett., vol. 9, p. 94, 1988.
[4] F. Assaderaghi, P. D. Ko, and C. Hu, “Observation of velocity overshoot
in silicon inversion layers,” IEEE Electron Device Lett., vol. 14, p. 484,
1993.
[5] Y. P. Tsividis, Operation and Modeling of the MOS Transistor. New
York: McGraw-Hill, 1987.
[6] J. G. Ruch, “Electron dynamics in short channels field-effect transistors,”
IEEE Trans. Electron Devices, vol. ED-19, pp. 652, 1972.
[7] G. Baccarani and M. R. Wordeman, “An investigation of steady-state
velocity overshoot in silicon,” Solid-State Electron., vol. 28, p. 407,
1985.
[8] K. Seeger, Semiconductor Physics: An Introduction, 5th ed. Berlin,
Springer-Verlag, 1991.
[9] E. Sangiorgi and M. R. Pinto, “A semi-empirical model of surfacescattering for Monte Carlo simulation of silicon N-MOSFET’s,” IEEE
Trans. Electron Devices, vol. 39, pp. 356, 1992.
846
IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 44, NO. 5, MAY 1997
[10] S. E. Laux and M. V. Fischetti, “Monte Carlo simulation of submicrometer Si N-MOSFET’s at 77 and 300 K,” IEEE Electron Device
Lett., vol. 9, p. 467, Sept. 1988.
[11] M. V. Fischetti and S. E. Laux, “Monte Carlo study of electron transport
in silicon inversion layer,” Phys. Rev. B, vol. 48, pp. 2244, 1993.
[12] T. Kobayasi and S. Kazuyuki, “Two-dimensional analysis of velocity
overshoot effects in ultrashort-channel Si MOSFET’s,” IEEE Trans.
Electron Devices, vol. ED-32, pp. 788–792, 1975.
[13] H. Ono, S. Masanobu, Y. Takashi, C. Fiegna, O. Tatsuya, and H. Iwai,
“A 40-nm gate-length n-MOSFET,” IEEE Trans. Electron Devices, vol.
42, pp. 1822–1830, 1995.
[14] M. V. Fischetti, S. E. Laux, and W. Lee, “Monte Carlo simulation of hotcarrier transport in real semiconductor devices,” Solid-State Electron.,
vol. 32, p. 1723, 1989.
[15] P. J. Price, “On the flow equation in device simulation,” J. Appl. Phys.,
vol. 63 p. 4718, 1988.
[16] K. K. Thornber, “Current equations for velocity overshoot,” IEEE
Electron Device Lett., vol. 3, p. 69, Mar. 1982.
[17] M. Artaki, “Hot-electron flow in an inhomogeneous field,” Appl. Phys.
Lett., vol. 52, p. 141, 1987.
[18] D. Chen, E. C. Kan, and U. Ravaioli, “An analytical formulation of
the length coefficient for the augmented drift-diffusion model including
velocity overshoot,” IEEE Trans. Electron Devices, vol. 38, p. 1484,
1991.
[19] M. R. Pinto, E. Sangiorgi, and J. Bude, “Silicon MOS transconductance
scaling into the overshoot regime,” IEEE Electron Device Lett., vol. 14,
p. 375, 1993.
[20] K. K. Thornber, “Relation of drift velocity to low-field mobility and
high-field saturation velocity,” J. Appl. Phys., vol. 51, pp. 2127, 1980.
[21] N. D. Arora, R. Rios, C. L. Huang, and K. Raol, “PCIM: A physically
based continuous short-channel IGFET model for circuit simulation,”
IEEE Trans. Electron Devices, vol. 41, p. 988, 1994.
[22] L. M. Dang and M. Konaka, “A two-dimensional computer analysis
of triode-like characteristics of short-channel MOSFETs,” IEEE Trans.
Electron Devices, vol. ED-27, pp. 1533–1539, 1980.
[23] N. Kotani and S. Kawazu, “Computer analysis of punch-through in
MOSFET’s,” Solid-State Electron., vol. 22, pp. 63–70, 1979.
[24] H. C. De Graaff and F. M. Klaassen, Compact Transistor Modeling For
Circuit Design. Berlin: Springer-Verlag, 1990, p. 176.
[25] F. Gamiz, J. A. Lopez-Villanueva, J. A. Jimenez-Tejada, I. Melchor, and
A. Palma, “A comprehensive model for Coulomb scattering in inversion
layers,” J. Appl. Phys., vol. 75, p. 924–934, 1994.
[26] J. Y.-C. Sun, M. R. Wordeman, and S. E. Laux, “On the accuracy
of channel length characterization of LDD MOSFET’s,” IEEE Trans.
Electron Devices, vol. ED-33, p. 1556, 1986.
[27] J. Chung, C. Jeng, G. May, P. K. Ko, and C. Hu, “Intrinsic transconductance extraction for deep-submicrometer MOSFET’s,” IEEE Trans.
Electron Devices, vol. 36, p. 140, 1989.
[28] J. A. Cooper, Jr. and D. F. Nelson, “High-field drift velocity of electrons
at the Si-SiO2 interface as determined by a time-of-flight technique,” J.
Appl. Phys., vol. 54, p. 1445, 1983.
[29] A. Modelli and S. Manzini, “High-field drift velocity of electrons in
silicon inversion layers,” Solid-State Electron., vol. 31, pp. 99, 1988.
[30] F. Fang and A. B. Fowler, “Hot electron effects and saturation velocities
in silicon inversion layers,” J. Appl. Phys., vol. 41, p. 1825, 1970.
[31] D. K. Ferry, “Transport of hot carriers in semiconductor quantized
inversion layers,” Solid-State Electron., vol. 21, p. 115, 1976.
J. B. Roldán received the degree in physics from
Granada University, Granada, Spain, in 1993.
Currently, he is a Teaching Assistant at the University of Granada. Since 1993, he has worked on
the MOS device physics including 2-D transport,
nonlocal effects, and Monte Carlo simulations. His
current interests are also related to SiGe and SiC
devices.
F. Gámiz received the degree in physics in 1991,
and the Ph.D. degree in 1994 from the University
of Granada, Granada, Spain.
Currently, he is an Associate Professor at the
University of Granada. Since 1991, he has been
working on the charge carriers in semiconductor
heterostructures. He has studied electron mobility in
silicon inversion layers by the Monte Carlo method.
His current research interests include the effects
of many-carriers on the electron mobility and the
theoretical interpretation of the influence of high
longitudinal electric fields on the electric properties of MOS transistors. His
current interests are also releated to SiGe, SiC, and SOI devices, and quantum
transport. He has coauthored several papers in all these subject areas.
J. A. López-Villanueva received the degree in
physics in 1984, and the Ph.D. degree in 1990
from the University of Granada, Granada, Spain. His
thesis was on the degradation of MOS structures by
Fowler–Nordheim tunneling.
Currently, he is an Associate Professor at the
University of Granada. Since 1985, he has been
working on deep-level characterization and mainly,
MOS device physics, including Fowler–Nordheim
and direct tunneling, quantum effects, 2-D transport,
effects of nonparabolicity, scattering mechanisms,
and Monte Carlo simulation of charge transport. He has coauthored several
papers in all these subject areas. His current research interest includes the
characterization, simulation, and modelling of electron devices, with emphasis
on the MOS transistor. His educational activities also include analog systems
for electronic instrumentation and power electronics.
J. E. Carceller received the degree in physics
in 1975, and the Ph.D. degree in 1979 from the
University of Barcelona, Barcelona, Spain.
He has been a Professor at the Universities of
Barcelona and Granada. He was engaged in the
research and characterization of deep levels in semiconductors. His current research interest includes
degradation of MOS structures and characterization of electron mobility in the channel of MOS
transistors.
Download