You are on page 1of 9

232 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 47, NO.

1, JANUARY 2000

On the Performance Limits for Si MOSFETs:


A Theoretical Study
Farzin Assad, Zhibin Ren, Dragica Vasileska, Member, IEEE, Supriyo Datta, Fellow, IEEE, and
Mark Lundstrom, Fellow, IEEE

AbstractPerformance limits of silicon MOSFETs are exam- Before we begin, we should ask whether the ballistic
ined by a simple analytical theory augmented by self-consistent limits discussed here are of concern for present-day devices.
SchrdingerPoisson simulations. The on-current, transconduc- Hot electrons near the drain have a mean-free-path of only
tance, and drain-to-source resistance in the ballistic limit (which
corresponds to the channel length approaching zero) are exam- a few Angstroms, much shorter than the channel length
ined. The ballistic transconductance in the limit that the oxide of present-day MOSFETs. But when the performance of
thickness approaches zero is also examined. The results show that present-day MOSFETs is compared to the corresponding
as the channel length approaches zero (which corresponds to the ballistic limits presented in this paper, we find that they operate
ballistic limit), the on-current and transconductance approach at roughly 3040% of the ballistic limit. The reason is that the
finite limiting values and the channel resistance approaches a
finite minimum value. The source velocity can be as high as about steady-state on-current is controlled by a very short region near
1.5 107 cm/s. The limiting on-current and transconductance the source [3]. Once the mean-free-path (which is longer near
are considerably higher than those deduced experimentally by a the source) is comparable to the length of this critical region,
previous study of MOSFETs with channel lengths greater than quasi-ballistic transport occurs. This is analogous to transport
0.2 m. At the same time, the transconductance to current ratio is in metal-semiconductor diodes, where the thermionic emission
substantially lower than that of a bipolar transistor.
(ballistic) theory applies if the mean-free-path is longer than the
Index TermsCharge carrier processes, MOSFETs, nanotech- distance over which the first of potential drop occurs
nology, semiconductor device modeling, semiconductor devices. (the so-called Bethe condition). Current levels that approach
the ballistic limit occur when the mean-free-path is comparable
I. INTRODUCTION to the length of the critical region of the device; it does not have
to be longer than the channel length itself.

A S silicon technology advances, questions about ultimate


performance limits arise. Toriumi et al. estimated perfor-
mance limits by extrapolating measured results to the limits of
In this paper, we examine the limiting performance of a
device corresponding to the nm node in the 1997
national technology roadmap for semiconductors (NTRS) [6].
zero channel length and zero oxide thickness [1]. Their results The device on-current, transconductance, and source-to-drain
suggested that the electron saturation velocity and finite inver-
resistance will be examined in two limits, the ballistic limit
sion layer capacitance set a limiting transconductance of about (corresponding to the channel length approaching zero or
3000 mS/mm. In this paper, we reexamine MOSFET limits by the mean-free-path approaching infinity) and the zero oxide
calculating the theoretical performance limits for an idealized, thickness limit. Although there are simplifying assumptions in
ballistic MOSFET using an approach similar to that of Natori the theoretical model, the results will indicate that the ultimate
[2] and our own recent work [3], [4]. For this study, we make MOSFET limits are considerably higher than previously esti-
use of self-consistent SchrdingerPoisson simulations [5] to mated and well above those currently achieved. On the other
examine the validity of several simplifying assumptions made hand, the transconductance limit is well below that projected
in earlier work. The results should be useful for assessing the by Johnson, who argued that as the channel length and oxide
performance potential of different transistors designs, for com- thickness approach zero, the transconductance of a MOSFET
paring measured device characteristics against ultimate limits, should approach that of a bipolar transistor [7]. If projections
and for identifying important technological and theoretical is- like those to be described in this paper are done for devices at
sues for silicon transistors at the end of the roadmap [6]. the end of the NTRS, they show that to achieve the performance
targets called for, devices will have to operate significantly
closer to their ultimate limits than todays devices. Parasitic
Manuscript received January 11, 1999; revised June 2, 1999. This work resistances and scattering at the oxide-silicon interface tend
was supported by the Semiconductor Research Corporation under Contract to increase as devices are scaled and will make it difficult to
98-SJ-089, the Defense Advanced Research Projects Administration Advanced
Microelectronics Program, and by the National Science Foundation Distributed
operate closer to the ballistic limits. New device designs are
Center for Advanced Microelectronics Simulations (DesCArtES). The review likely to be necessary to achieve device performance targets at
of this paper was arranged by Editor J. N. Hollenhorst. the end of the NTRS.
F. Assad, Z. Ren, S. Datta, and M. Lundstrom are with the School of Electrical
and Computer Engineering, Purdue University, West Lafayette, IN 47907-1285
USA (e-mail: lundstro@ecn.purdue.edu). II. THEORY
D. Vasileska is with the Department of Electrical Engineering, Arizona State
University, Tempe, AZ 85287-5706 USA. The ballistic current for a MOSFET is readily evaluated as
Publisher Item Identifier S 0018-9383(00)00169-6. described by Natori [2] and Datta [4]. A sketch of the derivation
00189383/00$10.00 2000 IEEE
ASSAD et al.: ON THE PERFORMANCE LIMITS FOR Si MOSFETS 233

statistics. For small drain biases, we find the drain-to-source


conductance as

(5)

The transconductance contributed by the th subband,

(6)

is obtained by differentiating the drain current. In the limit that


the oxide thickness approaches zero or its dielectric constant,
infinity, there is no voltage drop across the oxide, and the lim-
iting transconductance is

(7)

where is the surface potential.


Fig. 1. Assumed geometry of the ballistic MOSFET. Current flow is along the
x-direction, the confinement potential varies in the y -direction, and the width Equations (1)(7) summarize the important formulas, which
of the device is along the z -direction. are essentially those of Natori [2]. (A sketch of their derivation
is contained in the Appendix.) Our work extends that of
Natori by: 1) using self-consistent, one-dimensional (1-D)
is presented in the Appendix; the important results are summa-
SchrdingerPoisson simulations to treat all relevant subbands,
rized here. Fig. 1 shows the geometry. Current flow is in the
2) examining the limit in addition to the ballistic
-direction, the confining potential is in the -direction, and the
limit, and 3) making specific predictions for devices on the
width of the transistor is in the -direction. The drain current
NTRS [6]. The analysis procedure begins with a self-consistent
is sum of the contributions from each subband. The positive
SchrdingerPoisson simulation that determines the equilib-
states are populated by injection from the source, and the neg-
rium Fermi energy and the subband energies for a given gate
ative states by injection from the drain. The drain current
bias. The channel conductance can then be evaluated from (5).
contributed by subband, , is
To sweep out a common source characteristic, the inversion
layer density is fixed at its equilibrium value for the given gate
voltage, and (2) is solved for the Fermi energy at each drain
bias. At each drain bias, the corresponding current is evaluated
(1) from (1). Finally, by perturbing the gate bias, we also obtain
the transconductances by numerical differentiation.
where is the conductivity effective mass of subband, , and The SchrdingerPoisson solver used in this work has been
is the subband energy and is the FermiDirac integral described previously [5]; we summarize the essentials here.
of order one-half as defined by Blakemore [8]. Similarly, the The one-electron Schrdinger equation is solved self-con-
electron density in subband, , is sistently with Poissons equation with an effective potential
included to treat exchange and correlation. We use the density
functional theory of Hohenberg and Kohn [9], and Kohn and
Sham [10], and the interpolation formula developed by Hedin
(2) and Lundqvist, which is accurate over a large density range
[11]. We assume that the wavefunction goes to zero at the
where is the density of states effective mass for the sub- oxide-silicon interface. This assumption can be questioned for
band. Under high drain bias, injection from the drain is sup- very thin oxides, but does not appear to significantly affect the
pressed, and the on-current can be written as results presented here. The poly-silicon gates are modeled as
heavily-doped single-crystal silicon. Electrons in the poly-sil-
(3) icon and holes in the substrate are treated classically assuming
general FermiDirac statistics.
where Our 1-D treatment of MOS electrostatics implies a gradual
channel approximation at the source and drain ends of the
(4) channel. The source end is at the top of a potential energy
barrier, so the lateral electric field is zero and a 1-D treatment
is reasonable. (The second derivative is nonzero, however, so
is the product of the nondegenerate thermal velocity, two-dimensional (2-D) effects such as drain-induced barrier
, and a correction factor for FermiDirac lowering (DIBL) need to be accounted for separately.) At the
234 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 47, NO. 1, JANUARY 2000

drain end, the 1-D treatment can only be justified at low drain
voltages, but when the voltage exceeds a few , carrier
injection from the drain is small, so the assumption has little
effect on the drain current.

III. RESULTS
The device examined was taken from the nm node
of the national technology roadmap for semiconductors [6]. The
oxide thickness is 1.5 nm, and the substrate is uniformly doped
at cm-3. This substrate doping results in a
threshold voltage of about 0.4 V, which is much too high for the
assumed 1.0 V power supply. In practice, however, this value
would be lowered by 2-D, short-channel effects which are not
considered here. Results are quoted, therefore, at a gate over-
drive of V, which would result for a more appro-
Fig. 2. Inversion layer density versus the gate voltage as obtained by a
priate of 0.2 V. [Note that the uniform channel doping was self-consistent solution to the Schrdinger and Poisson equations. Two cases
assumed only to provide a baseline against which more sophis- +
are shown: 1) the equilibrium solution when both k and k states 0
ticated channel doping profiles can be compared. For example, +
occupied, solid lines and 2) the ballistic solution for which only the k states
are occupied, dashed lines. Solutions for both a metal gate and for a polysilicon
it would be interesting to compare the ultimate performance of a gate are shown. For comparison, a classical calculation for a metal gate
super steep retrograde (SSR) channel profile [12] against a uni- + 0
transistor with both k and k states occupied is also shown (dotted line).
form profile.] Unless otherwise stated, all simulations are for
K. In the remainder of this section, we present several
results and discuss some issues. At the end of the section, we very nearly the same. This occurs because of charge balance in
summarize the ultimate performance parameters for this model the MOS system. As the negative half is suppressed by the in-
device. creasing drain bias, the surface potential increases, pushing the
In equilibrium, the positive and negative states in the Fermi level deeper into the conduction band, which increases
carrier distribution function are both occupied. In a ballistic and maintains charge balance. This is a kind of DIBL effect
MOSFET, however, positive states are populated by injec- (drain induced barrier lowering) but it is not due to 2-D electro-
tion from the source, and the negative states by injection statics but, rather, to the need to maintain charge balance in the
from the drain. If the drain bias exceeds several , 1-D electrostatics as the distribution function is distorted by the
therefore, the carrier distribution function at the source will drain bias. There is only a small decrease in , which can be
consist of only positive states. The effects of this strongly interpreted as a small reduction in the effective gate capacitance
off-equilibrium distribution on MOS electrostatics will be for the ballistic MOSFET. We conclude that when computing a
examined next. common source characteristic, fixing the value of at its equi-
Fig. 2 shows the computed inversion layer density versus gate librium value as varies, is a good approximation.
voltage as obtained by self-consistent, SchrdingerPoisson Fig. 3(a) shows the common source characteristics of the
simulations [5]. Results for both a metal gate and a polysilicon MOSFET as computed using (1). For this computation, we eval-
gate are shown. The metal gate is taken to be aluminum (with uated in equilibrium for a given gate voltage, then fixed at
a workfunction difference of eV) while the the equilibrium value as was swept from zero to the supply
polysilicon gate is doped to 1 1020 cm-3. As expected, the voltage. The ballistic drain current is seen to saturate at a drain
equilibrium versus characteristics (solid lines) are well bias of a few , which is much lower than the saturation
fit by voltage of a typical MOSFET. Beyond the saturation voltage,
the drain current is constant when 2-D electrostatics, which pro-
(8) duces DIBL in a real device, are not treated. (Also shown in
Fig. 3(b) is a common source characteristic showing the effects
where is the total gate capacitance. For the metal gate re- of DIBL, which effectively lowers , or raises , as
sults, the capacitance extracted from the slope of versus increases.) Finally, the characteristics clearly shows
gives a capacitance about 25% less than the oxide capacitance the detrimental effect of the polysilicon depletion on the drive
due to well-known carrier degeneracy and quantum mechanical current.
confinement effects [5]. Polysilicon depletion further reduces In Fig. 4, we plot the ballistic characteristic for the
the gate capacitance. metal gate device at a supply voltage of
To examine the effect of having only the states occu- V. The drain current is proportional to inversion layer density
pied, which will occur in a ballistic MOSFET under a drain times the thermal injection velocity. The inversion layer density
bias of a few or greater, we reduced the 2-D density of is proportional to ( ) as given by (8). The thermal in-
states by a factor of two and performed the simulation again jection velocity (4) is constant for nondegenerate conditions and
(dashed lines in Fig. 2). The results show that even when only proportional to when fully degenerate. Conse-
one-half of the states are occupied, the inversion layer density is quently, we expect to vary linearly with ( ) under
ASSAD et al.: ON THE PERFORMANCE LIMITS FOR Si MOSFETS 235

(a)
Fig. 4. I versus V for the ballistic MOSFET at a drain voltage of 1.0 V.

(b)

Fig. 3. (a) Common source characteristics of metal gate (dotted lines) and
Fig. 5. The ratios, n =n and I =I versus V for the metal gate,
polysilicon gate (dashed lines) ballistic MOSFETs. (b) Common source
ballistic MOSFET.
characteristics of a metal gate ballistic MOSFET showing the effect of DIBL.
Without DIBL (solid line) and with an assumed DIBL of 100 mV/V (dashed
line).
If one describes the on-current by

(9)
nondegenerate conditions and as under degen-
erate conditions. Fig. 4 indicates that the device is operating be- then the upper limit for the average velocity at the source,
tween these two limits. , is the thermal injection velocity. In practice, scattering
Natoris analytical treatment assumed that only one subband will reduce below the thermal injection velocity [3]. In
was occupied [2]. Fig. 5 shows that about 90% of the inver- small transistors, strong velocity overshoot occurs within the
sion layer density and drain current arises from electrons in the channel. Through its influence on the self-consistent channel
first subband. The fraction increases slightly as the gate voltage potential, velocity overshoot affects the electric field at the
increases, because quantum confinement increases the energy source, and therefore carrier backscattering and consequently
spacing to the second subband, then decreases for high gate volt- [3]. But velocity overshoot does not change the upper
ages as the first subband becomes degenerate. For this device, limit for .
the one-subband approximation is adequate. Fig. 6(a) shows the The transconductance of the ballistic MOSFET is
thermal injection velocity versus gate voltage as obtained from
the SchrdingerPoisson simulation and (4). The thermal in-
jection velocity is constant only in the nondegenerate regime, (10)
which occurs below threshold. Note that under on-current con-
ditions, the thermal injection velocity approaches 2 107 cm/s, If the second term is small, then the transconductance is a mea-
as has been noted by Natori [2]. In Fig. 6(b), we compare the sure of the thermal injection velocity. Fig. 7 compares the com-
thermal injection velocity to its degenerate limit. For this de- puted transconductance with that obtained from the first term on
vice, fully degenerate conditions do not occur. the right-hand side of (10). In the subthreshold region, where the
236 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 47, NO. 1, JANUARY 2000

(a) (a)

(b) (b)

Fig. 6. (a) Ballistic injection velocity at the source versus gate voltage, V , Fig. 8. (a) Ratio of transconductance divided by drain current, versus
for the metal gate, ballistic MOSFET. Also shown is the injection velocity for V for the metal gate, ballistic MOSFET. (b) The ratio of the intrinsic
the first subband alone. (b) Ratio of the degenerate limit velocity to the actual transconductance, g^ , to the drain current versus surface potential.
injection velocity for electrons in the lowest subband versus V .

velocity from the transconductance by ignoring the second term,


as is common practice, the result would be higher than the ac-
tual thermal injection velocity.
A useful way to assess the transconductance is to ex-
amine the ratio, . For a bipolar transistor, this ratio is
. It has been argued that for a MOSFET with the
oxide thickness approaching zero, should approach
the bipolar limit [7]. Fig. 8(a) plots versus for
the model transistor. Below threshold, the transconductance
does approach the bipolar limit. (It is a little lower than the
bipolar limit because the subthreshold slope of the MOSFET is
higher than the theoretical minimum of 60 mV/decade.) Above
threshold, however, is only about 5% of the bipolar
limiteven for the assumed oxide thickness of 1.5 nm. In the
limit that approaches zero, we can evaluate the limiting
Fig. 7. Transconductance versus V for the ballistic MOSFETs
transconductance, from (7). The result is shown in Fig. 8(b).
(solid lines). Also shown is the product of C V (dashed lines). The Below threshold, is even closer to the bipolar limit. (It
transconductance is seen to be higher than C V for V > 0:6 V. is a little lower because of the finite inversion layer thickness.)
Above threshold, the ratio drops, but it is a sizable fraction
thermal injection velocity is constant, the second term in (10) is of the bipolar limit. The limiting transconductance increases
negligible, but above threshold it is not. If one were to deduce a with surface potential; its maximum value will be set by the
ASSAD et al.: ON THE PERFORMANCE LIMITS FOR Si MOSFETS 237

TABLE I
COMPUTED PERFORMANCE OF THE MODEL BALLISTIC

DEVICE AT 300K. THE THRESHOLD VOLTAGE WAS 0:4 V, THE GATE
OVERDRIVE WAS (V 0 V ) 0:8 V, AND THE DRAIN VOLTAGE WAS 1.0 V

TABLE II
COMPUTED PERFORMANCE OF THE MODEL BALLISTIC DEVICE AT 77K. THE

THRESHOLD VOLTAGE WAS 0:5 V AND THE GATE OVERDRIVE WAS
(V 0
V ) 
0 :8 V

Fig. 9. Intrinsic drain to source resistance, R , versus V for the metal gate,
ballistic MOSFET.

maximum surface potential that can be achieved. Finally, note


that these results show that a hypothetical MOSFET with a
gate insulator having an infinite dielectric constant (so that its
times higher than the on-current target. Current-day devices op-
electrical thickness is zero) and an infinite bandgap (so that
erate at about the same percentage of the ballistic limit, so it ap-
there is no penetration of the wavefunction into the insulator),
pears that scaling current devices to the 100 nm node should be
still delivers a lower transconductance than a corresponding
capable of producing the required on-current. If one looks fur-
bipolar transistor. In practice, tunneling through the gate
ther ahead, however, the gap between the ballistic limit and the
insulator would further reduce performance, so that MOSFETs
performance target narrows. For example, for a model
will always deliver considerably lower transconductance that
nm devices with a 1.0 nm gate oxide and a power supply of 0.6
bipolar transistors.
V, the ballistic current is only 1.5 times the target on-current.
Finally, we examine the intrinsic drain to source resistance of
This suggests that to meet the on-current targets with low power
the ballistic MOSFET. Conventional MOSFET theory predicts
supply voltages, new device architectures might be needed.
a channel resistance that goes as the length of the channel, so
Operation at liquid nitrogen temperatures improves the per-
in the ballistic limit, where L approaches zero, there should be
formance of MOSFETs; typically, the on-current increases by
no channel resistance. Equation (5), however, shows that the
about 30% [14]. In the ballistic MOSFET, the on-current is de-
ballistic MOSFET has a finite resistance. This resistance, due
termined by the thermal injection velocity, which may be ex-
to the finite number of transverse modes available to conduct
pected to decrease at low temperatures. Table II shows the com-
current, is analogous to the quantum contact resistance
puted device performance parameters at K. Note that
of a mesoscopic device [13]. It sets a lower limit for the resis-
the on-current improves at low temperatures. The reason may
tance of a MOSFET, independent of contact technology. Fig. 9
be seen from Fig. 10, which compares the thermal injection ve-
is a plot of the intrinsic resistance of the ballistic MOSFET.
locity versus gate voltage at K and 77K. At low gate
This gate-voltage dependent intrinsic resistance is large enough
voltages, the inversion layer is nondegenerate, and the carrier
that it may have to be accounted for in analyzing experiments.
velocity is considerably lower at 77K. Above threshold, how-
A simple physical explanation for the finite resistance of the
ever, carrier degeneracy, which raises , is stronger at low
ballistic channel is that the net drain to source current is the
temperatures, so the injection velocity under maximum gate bias
difference between the current thermionically injected from
is actually higher at 77K than at 300K. Real devices dont op-
the source and that thermionically injected from the drain.
erate at the ballistic limit, so the reduction in scattering at low
The current injected from the drain is lower by a factor of
temperatures further boosts the on-current. A quantitative anal-
in the nondegenerate limit. At small drain
ysis of the magnitude of the effect would also have to consider
biases, the exponential can be expanded, and we find that the
possible changes in the saturation drain voltage and in the par-
net current is proportional to the drain voltage, which gives a
asitic resistances.
finite resistance.
In practice, device performance will fall well below the bal-
The computed performance limits for the nm bal-
listic limits projected here. Series resistance will become in-
listic MOSFET are summarized in Table I.
creasingly important as the channel resistance decreases. Scat-
tering by phonons and at the oxide-silicon interface needs to
IV. DISCUSSION
be accommodated in a realistic theory. We have also neglected
The performance limits for the ballistic MOSFET presented quantum mechanical reflections within the channel and from
in Table I are comfortably above the performance targets in the the drain and source, which may become significant as devices
NTRS [6]. For example, the ballistic on-current is almost four shrink. Our use of 1-D SchrdingerPossion simulations for
238 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 47, NO. 1, JANUARY 2000

much smaller, about 1.1 nm. Note that the limiting transcon-
ductance, , and the effective inversion layer thickness, , both
depend on the surface potential. The lower and larger ob-
served by Toriumi et al.suggest that the experimental data were
taken at lower surface potentials than assumed in the theoret-
ical calculations. Series resistance causes the transconductance
to peak at lower surface potentials and may explain the discrep-
ancy between the theory and experiment.

V. SUMMARY
We presented a simple procedure for estimating the per-
formance limits of silicon MOSFETs as channel lengths and
oxide thicknesses approach zero. In addition to the ballistic
limit current, the procedure also gives the ballistic limit
transconductance and the channel resistance. The procedure
Fig. 10. The injection velocity versus gate voltage for the model device. Solid assumed a gradual channel approximation at the source end of
line, T = 300K. Dashed line, T = 77 .K the channel and then made use of 1-D SchrdingerPoisson
simulations of the MOS problem. (This approach misses
MOS electrostatics needs to be examined by 2-D simulations important 2-D effects such as DIBL, which have to be added
of the same problem. Serious comparisons of the ballistic limit in separately.) We also treated the channel and drain as perfect
results with measured nanoscale device characteristics such as absorbers for electrons injected from the source, which neglects
those reported in [14] may also help us understand the signif- quantum mechanical reflections that might occur in very small
icance of such effects. This work is a prelude to that kind of transistors. These issues should be addressed by more sophisti-
serious comparison with experiment. It provides a clear, simple cated simulations, but the approach presented here provides a
procedure to estimating upper limits, so that the performance clear conceptual way to think about transistor limits, and the
of current devices and potential replacements can be gauged simplifying assumptions are physically based and will provide,
against performance limits. we believe, reasonable upper limits for device performance.
The theoretical results reported here should be compared to In agreement with a previous experimental study, we found
those of Toriumi et al.[1] who conducted a careful experimental that as the channel length approaches zero (or as the device
study to establish performance limits by extrapolating measured becomes ballistic) the drain current and transconductance have
results to zero channel length (which corresponds to our ballistic a finite limit. We also found a finite channel resistance, even
assumption) and zero oxide thickness. Toriumi et al. examined when the channel is ballistic. We found that the carrier velocity
the transconductance, at the source end of the channel [ in (9)] can be as high
as cm/s in the ballistic limit. The MOSFET perfor-
(11) mance limits established here are well above those of current
day devices and considerably higher than those deduced by To-
riumi et al. [1]. For example, the drive current targeted for the
where is the transconductance in the limit of zero oxide nm node of the NTRS is only about one fourth of the
thickness as defined by (7). Using ballistic limit current. Even for a 1.5-nm thin oxide, however,
the transconductance to current ratio is only about 5% of the
(12) value for a bipolar transistor. As one looks further ahead, how-
ever, it appears that devices will have to operate significantly
where is the semiconductor capacitance, (11) can be expressed closer to the ballistic limit if the NTRS targets are to be met.
as The approach presented here should be useful in assessing new
device designs as well as in understanding the performance lim-
(13) iting factors of present-day devices.

We have defined the effective thickness of the inversion layer APPENDIX


by . According to (13), a plot of inverse transcon- When the SiO2/Si interface is parallel to the [100] plane, the
ductance versus oxide thickness is linear if the effective thick- six equivalent minima of the bulk silicon conduction band split
ness of the inversion layer is constant. Toriumi et al.observed into two sets of subbands [16]. The first set consists of the two
this linear dependence experimentally [1]. They found a lim- equivalent valleys with in-plane effective density-of-states mass
iting transconductance of mS/mm, which is significantly (where a factor of two accounts for the valley de-
lower than the value we obtained. From the intercept of their generacy) and perpendicular effective mass of . The second
plot, an effective inversion layer thickness of nm can set ( -band) consists of the four equivalent valleys with
also be inferred. The effective inversion layer thickness obtained (again, factor of four accounts for valley degeneracy)
from our self-consistent, SchrdingerPoisson simulation was and the perpendicular effective mass is . The energy levels
ASSAD et al.: ON THE PERFORMANCE LIMITS FOR Si MOSFETS 239

associated with the first set comprise the so-called unprimed The results of our study show that this assumption is, for the
ladder of subbands, whereas those associated with the second most part acceptable. In this case, the total carrier concentration
set comprise the primed ladder of subbands. is the first subband population, so
The drain current of a MOSFET is the sum of the contri-
butions from the unprimed and primed subbands. The ballistic (10)
limit current contribution from each subband is computed from
and we may, therefore, solve (2) for and, therefore
(A1) perform the entire analysis analytically. A SchrdingerPoisson
simulation may be needed to specify and .
It is useful to examine the one-subband analytical results in
where is a normalization area, the degenerate limit. In this case

(A2)
(A11)
the Fermi function for electrons injected from the left contact,
Notice that (A11) is the standard, equilibrium result for single
and
spherical valley; it is actually a factor of two smaller because
(A3) only states are occupied, but then it is multiplied by two to
account for valley degeneracy. Using the degenerate limit of the
for electrons injected from the right contact. Parabolic energy FermiDirac integral
bands were assumed, and we assumed that the MOSFET width,
, was large so that the sum over could be replaced by an (12)
integral (see [17], Sections 1.2.3 and 3.1) for a discussion of how
such integrals are worked out). The result of evaluating (A1) in the degenerate thermal injection velocity becomes
this manner gives the drain current carried by each subband, (1).
In the drain current expressions, one must use the proper (13)
density of states and conductivity effective masses. For the un-
primed ladder Note that in the one-subband approximation, the thermal injec-
(A4) tion velocity is independent of temperature, if a fixed gate over-
drive is maintained. Fig. 10 shows that the injection velocity at
and for the primed ladder high gate voltages is higher at 77K than at 300K. This occurs
because at 300K, upper subbands are partially occupied, which
(A5) lowers the average injection velocity somewhat.
The on-current is now readily evaluated from the product of
Also
and to find
(A6)
on
for the unprimed ladder and

(A7) (A14)

for the primed ladder. The carrier density for each subband is Finally, we can find the one-subband, channel conductance in
given by (2), and the total inversion layer density per unit area the degenerate limit as
is the sum of the carrier densities of each subband. Note that
the factor of two in the denominator of (2) arises because in (A15)
the ballistic limit, only the states are occupied when is
greater than a few . Equation (A15) can be expressed in a more familiar form if we
To evaluate the source to drain resistance in the ballistic limit, write it in terms of the number of transverse modes at the Fermi
note that for small drain-to-source voltages energy

(A8) (A16)

and the source to drain channel conductance becomes In (A16), the factor of two comes from valley degeneracy. Using
(A16) in (A15), we find
(A9)
(A17)
which can be evaluated to find (5).
A common assumption in the analysis of MOS problems is to where , etc., because of the valley degeneracy.
assume that only a single subband is occupied (e.g., Natori [2]). Equation (A17) is a familiar result in mesoscopic physics [9]. In
240 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 47, NO. 1, JANUARY 2000

the degenerate limit, each transverse mode contributes to Zhibin Ren was born on October 2, 1969 in
the ballistic conductance. Zhengzhou, China. He received the B.S. degree
from Zhejiang University, China, and the M.S.
degree from the University of Massachusetts,
Dartmouth, in 1991 and 1997, respectively, all in
REFERENCES physics. Currently, he is a Ph.D. candidate in the
[1] A. Toriumi, M. Iwase, and M. Yoshimi, On the performance limit for Department of Electrical Computer Engineering,
Si MOSFET: Experimental study, IEEE Trans. Electron Devices, vol. Purdue University, West Lafayette, IN. His work
35, pp. 9991003, 1988. includes research on deep submircon device physics
[2] K. Natori, Ballistic metal-oxide-semiconductor field effect transistor, and bipolar and MOS transistor design. He has been
J. Appl. Phys., vol. 76, pp. 48794890, 1994. engaged in 25-nm generation transistor research on
[3] M. S. Lundstrom, Elementary scattering theory of the MOSFET, IEEE at Purdue since 1997.
Electron Device Lett., vol. 18, pp. 361363, 1997.
[4] S. Datta, F. Assad, and M. S. Lundstrom, The Si MOSFET from a
transmission viewpoint, Superlattices and Microstructures, vol. 23, pp. Dragica Vasileska (M97) received the B.S.E.E.
771780, 1998. (diploma) and the M.S.E.E. degrees from the
[5] D. Vasileska, D. K. Schroder, and D. K. Ferry, Scaled silicon University of Cyril and Methodius, Skopje, Republic
MOSFETs: Degradation of the total gate capacitance, IEEE Trans. of Macedonia, in 1985 and 1992, respectively, and
Electron Devices, vol. 44, pp. 584587, 1997. the Ph.D. degree from Arizona State University
[6] National Technology Roadmap for Semiconductors: Semiconductor In- (ASU), Tempe, in 1995.
dustry Assoc., 1997. From 1995 until 1997, she held a faculty research
[7] E. O. Johnson, The insulated-gate field-effect transistorA bipolar position within the Center of Solid State Electronics
transistor in disguise, RCA Review, vol. 34, pp. 8094, 1973. Research at ASU. In the fall of 1997, she joined
[8] J. S. Blakemore, Approximations for the FermiDirac integrals, espe- the faculty of Electrical Engineering at ASU. Her
cially the function, [ ], used to describe electron density in a semi-
research interests include semiconductor device
conductor, Solid-State Electron., vol. 25, pp. 10671076, 1982. physics and semiconductor device modeling, with strong emphasis on quantum
[9] P. Hohenberg and W. Kohn, Inhomogeneous electron gas, Phys. Rev., transport and Monte Carlo particle-based simulations. She has published 34
vol. 136, pp. B864B871, 1964. journal publications, 16 conference proceedings refereed papers, and has given
[10] W. Kohn and L. J. Sham, Self-consistent equations including exchange several invited talks.
and correlation effects, Phys. Rev., vol. 140, pp. A1133A1140, 1965. Dr. Vasileska is a member of the APS.
[11] L. Hedin and B. I. Lundqvist, Explicit local exchange-correlation po-
tentials, J. Phys. C: Solid State Phys., vol. 4, pp. 20642083, 1971.
[12] G. G. Shahidi, B. Davari, T. J. Bucelot, P. A. Ronsheim, P. J. Coane,
S. Pollack, C. R. Blair, B. Clark, and H. H. Hansen, Indium channel Supriyo Datta (M82SM93F96) received
implant for improved short-channel behavior of submicrometer the B.Tech degree from the Indian Institute of
NMOSFETs, IEEE Electron Device Lett., vol. 14, pp. 409411, 1993. Technology, Kharagpu, in 1975 and the Ph.D. degree
[13] S. Datta, Electronic Transport in Mesoscopic Systems. Cambridge, from the University of Illinois at Urbana-Champaign
U.K.: Cambridge Univ. Press, 1997. in 1979.
[14] G. A. Sai-Halasz, M. R. Wordeman, D. P. Kern, E. Ganin, S. Kishton, In 1981, he joined Purdue University, where he is
D. S. Zicherman, H. Schmid, M. R. Polcari, H. Y. Ng, P. J. Restle, T. H. currently a Distinguished Professor in the School of
P. Chang, and R. H. Dennard, Design and experimental technology for Electrical and Computer Engineering. He has written
0.1 m gate length, low-temperature operation FETs, IEEE Electron three books: Surface Acoustic Wave Devices (En-
Device Lett., vol. EDL-8, pp. 463467, 1987. glewood Cliffs, NJ: Prentice-Hall, 1986), Quantum
[15] G. Timp, A. Agarwal, F. H. Baumann, T. Boone, M. Buonanno, R. Phenomena (Boston, MA: AddisonWesley, 1989),
Cirelli, V. Donnelly, M. Foad, D. Grant, M. Green, H. Grossmann, and Electronic Transport in Mesoscopic Systems (Cambridge, U.K.: Cambridge
S. Hillenius, J. Jackson, D. Jacobson, R. Kleiman, A. Kornblit, F. Univ. Press, 1995), and has coauthored more than 100 journal articles. His
Klemens, J. T. C. Lee, W. Mansfield, S. Moccio, A. Murrell, M. O. current research interests are generally centered around the physics of nanos-
Malley, J. Rosamilia, J. Sapjeta, P. Silverman, T. Sorsch, W. W. Tai, D. tructures and includes molecular electronics, spin electronics, and mesoscopic
Tennant, H. Vuong, and B. Weir, Low leakage, ultra-thin gate oxides superconductiveity in addition to nanoscal device physics.
for extremely high performance sub-100 nm nMOSFETs, in IEDM Dr. Datta received a Presidential Young Investigator Award and an IEEE Cen-
Tech. Dig., Washington, DC, Dec. 1997. tennial Key to the Future Award in 1984 and the Frederick Emmons Terman
[16] T. Ando, A. B. Fowler, and F. Stern, Electronic properties of two-di- Award from the ASEE in 1994. He is a fellow of the APS.
mensional systems, Rev. Mod. Phys., vol. 54, pp. 437672, 1982.
[17] M. Lundstrom, Fundamentals of Carrier Transport. Reading, MA:
Addison-Wesley, 1990.
Mark Lundstrom (SM80F94) received the
B.E.E. and M.S.E.E. degrees from the University of
Minnesota, Minneapolis, and the Ph.D. degree from
Purdue University, West Lafayette, IN, in 1973,
1974, and 1980, respectively.
From 1974 to 1977, he was with Hewlett Packard,
Farzin Assad was born in Tehran, Iran. He was where he worked on the development of second-gen-
received the B.S.E.E. degree with honors and the eration NMOS integrated circuit process. He is cur-
M.S.E.E. degree in 1990 and 1991, both from the rently a Professor of electrical engineering at Purdue
University of Minnesota, Minneapolis. In 1995, University, where he has also served as the Director
he joined the device simulation team at Purdue of the Optoelectronics Research Center and as an As-
University, West Lafayette, IN, where he is currently sistant Dean of Engineering. His research interests center on the physics of semi-
pursuing the Ph.D. degree. His research interests are conductor devices and include carrier transport, ultrasmall transistors, and com-
in the area of transport in ultra-small devices and putational electronics. He is the author of a textbook, Fundamentals of Carrier
electromagnetics. Transport (Boston, MA: Addison-Wesley).
Dr. Lundstrom is the 1993 recipient of the ASEE Frederick Emmons Terman
Award.

You might also like