CMOS Scaling
CMOS Scaling
CMOS Scaling
I. INTRODUCTION
After more than two decades of relentless scaling to ever
smaller dimensions for higher packing density, faster circuit
speed, and lower power dissipation, CMOS technology
has become the prevailing technology for very large scale
integration (VLSI) applications today. These advances led
to computers and networks with far superior performance
and dramatically reduced cost per function. Currently, 0.35m CMOS technology with 0.25- m channel length is
being used in the manufacturing of 64-Mb DRAMs and
200-MHz microprocessors with the number of transistors
per chip in the
(fT)
of
n-
and
Optical photolithography has exceeded previously predicted resolution limits many times over by a combination
of improved lenses with higher numerical aperture and
the use of shorter wavelength illumination. Todays most
advanced production lithography equipment uses excimer
laser sources at a wavelength of 248 nm (KrF), offering a
resolution near 0.25 m. Further improvements in resolution may be attained at the 193-nm wavelength (ArF). A
prototype 193-nm stepper has already shown lithographic
resolution to 0.18 m [3]. Resolution enhancement tech487
Fig. 6.
80-nm-wide lines patterned in APEX-E resist by X-ray
lithography. The mask-wafer gap was 25 m.
niques, such as phase shifting, are capable of imaging features in the 0.100.12- m range, with a 248-nm source [4].
This technique relies on material or topographic changes
on the optical mask to vary the phase of the illuminating
radiation. The resulting interference effectively sharpens the
image at the wafer plane. Because this effect is geometry
dependent, the phase-shifting technique has thus far not
been demonstrated to be generally applicable to arbitrary
device geometries which may be encountered in a chip
design. With the exception of near-field techniques, which
may be impractical for device fabrication applications,
there are no current expectations that optical lithographic
techniques will extend into the sub-100-nm regime. For the
fabrication of such ultrasmall devices, optical lithography
may be used for noncritical levels, in a mix-and-match
scheme, where the critical features are defined by electron
beam lithography or X-ray lithography.
X-ray lithography is a prime candidate for high-resolution
patterning of critical features for sub-100-nm CMOS applications. Fig. 6 shows line/space features with a resolution
80 nm patterned by synchrotron X-ray lithography with a
mask-wafer gap of 25 m. Chen et al. have demonstrated
50-nm features patterned by proximity X-ray lithography
with a 5- m gap [5]. In a near-contact printing mode, 30nm features have been resolved by X-ray lithography [6],
which is sufficient to pattern gates for MOSFET devices
near the currently perceived limits of operation (see later
discussions). The challenges in implementing an X-ray
lithography technology lie primarily in mask fabrication. Xray masks consist of thin ( 25 m) membranes of Si or a
Si compound such as Si N or SiC, patterned with an X-ray
absorbing material. Electroplated gold has been used as the
absorber material. Precise control of mechanical stress in
the absorber-covered membrane must be maintained, as it
has a direct effect on the image placement accuracy in X-ray
lithography. In addition, since X-ray proximity printing is a
replication, more stringent control of defects is required
as compared to systems using image reducing optics.
Another form of X-ray lithography under investigation
for sub-100 nm applications has come to be known as
extreme ultraviolet (EUV) lithography [7]. This technique
488
where
is the total node capacitance being switched
(either up or down) in a clock cycle and is the clock frequency. As CMOS technology advances, clock frequency
goes up. The total switching capacitance is likely to increase
as well, as one tries to integrate more circuits into the
same or even larger chip area. The active power of todays
microprocessors is already in the 1020-W range. Besides
power management systems with architectural innovation,
the most effective way to curb the growth of active power is
to reduce the power supply voltage. Fig. 8 plots the power
per device of a 0.1- m CMOS ring oscillator versus the
gate delay by varying the power supply voltage [16]. Also
shown are two points representing similar circuits fabricated
with the current 3.3-V CMOS technology and with the
upcoming 2.5-V technology. It is clear that by reducing the
power supply voltage, one achieves a dramatic reduction
TAUR et al.: CMOS SCALING INTO THE NANOMETER REGIME
where
is the total turnedoff device width with
across it,
is the worst case off-current per device
width at 100 C,
is the extrapolated current per width at
threshold voltage (of the order of 110 A/ m for 0.1- m
devices),
is a dimensionless factor normally
(see
Section III-C), and
is the worst case threshold voltage
at 100 C. Typically,
is lower than the nominal roomtemperature threshold voltage, , shown in Fig. 1 by about
200 mV because of both the temperature difference (25100
C) and the threshold uncertainties due to channel-length
tolerances. The latter stems from the short-channel effect to
be discussed in the next section. The primary reason that
cannot be scaled is because the inverse subthreshold slope,
(ln10)(
), a measure of the transistor turn-off rate
versus gate voltage, is largely driven by thermally activated
diffusion and is independent of power supply voltage and
channel length. In fact, even if
is kept constant, the
leakage current of turnedoff devices would increase in
1
proportion to
and
. The off-state leakage
current would further increase by about 10 for every 0.1V reduction of . Fortunately, the leakage current and
therefore the standby power is quite low in todays 3.3-V
chips. There is also a limited reduction of short-channel
tolerances with lower power supply voltages. These
allow some room for a slightly downward trend of ,
as shown in Fig. 1. For room-temperature CMOS devices,
1 The current at threshold condition I is proportional to the inversion
0
charge density at threshold, Qi
(12)(kT=q )Cox , where Cox =
"ox =tox is the gate oxide capacitance per unit area.
489
proportion to . The latter requires increased channeldoping concentration which, for a uniformly doped channel,
leads to higher depletion charge and electric field at the
silicon surface. These in turn cause the potential across
the oxide and therefore the threshold voltage to go up. To
reduce the gate-controlled depletion width while fulfilling
the
-reduction trend depicted in Fig. 1, a retrograde,
i.e., lowhigh, channel doping is needed below 0.2- m
channel length [1]. Fig. 10 shows a schematic band-bending
diagram at the threshold condition of an extreme retrograde
profile with an undoped surface layer of thickness . For
the same gate depletion width
, the surface electric
field
and the total depletion charge of an extreme
retrograde channel is one-half of that of a uniformly doped
channel. This reduces threshold voltage and improves mobility.
Retrograde channel doping represents a vertically
nonuniform profile that allows the threshold voltage
to be decoupled from the gate-controlled depletion
width. However, the body-effect coefficient,2
, and the inverse
subthreshold slope, (ln10)(
), are still coupled to
the gate depletion width
. For a given
, reduction
in
improves short-channel effect but compromises
substrate sensitivity and subthreshold slope. Halo doping or
nonuniform channel profile in the lateral direction provides
yet another degree of freedom which can be tailored
to further minimize
-tolerances due to short-channel
effect [1]. Fig. 11 shows schematically an idealized, 2-D
nonuniform channel-doping profile. Pockets of high-doping
regions are placed at two lower corners of the gatecontrolled depletion region where the potential difference
(band bending) between the source/drain and the substrate
is the highest. These regions are partly depleted by the
2 The
m01, is usually referred to as the body effect. Body effect tends to degrade
Fig. 12. High-resolution cross-sectional TEM image of polysili oxide. The cross section is
con gate over a thermally grown 33-A
in h110i plane of silicon. The silicon crystal lattice spacing (3.13
provides a calibration for oxide thickness.
A)
characteristics for
(
A).
are also shown. At a gate bias of 1.5 V,
oxides down to 15 A
the current density increases by ten orders of magnitude as
The current
the oxide thickness decreases from 36 to 15 A.
Fig. 14. Calculated MOS capacitance under inversion normalized to the oxide capacitance as a function of gate oxide thickness. The dashed curves are calculated from a classical model
with Fermi-Dirac statistics; the solid curves are calculated from
the quantum mechanical model. The differences between the
metal-gate and the polysilicon-gate (doped to 1020 cm03 concentration) cases are due to polysilicon depletion effects.
G. Interconnect Delays
If the performance of processors is to keep pace with
the speed improvement of devices, special attention has
to be paid to interconnections. This is because the delay
stemming from wire resistance, commonly referred to as
delay, to first order is
(8)
and has to be included with other delay components. Here,
and
are the resistance and capacitance of unit wire
length,
is the wire length, and
is the load at
the end of the wire. The part in the
delay due to
the wire alone does not decrease in spite of scaling to
smaller dimensions. The factor that improves the
term through shorter
is negated by the increase in
due to wire cross-sectional shrinkage. Wire capacitance
in the meantime remains constant, around 0.2 pF/mm for
minimum width wires with oxide dielectric [33].
Traditionally the wire
delays were barely noticeable even on long wires, but for high-performance CMOS
CPUs, the resistance in the wiring can be critical. If
one looks carefully at roles various interconnects play, a
performance-oriented approach immediately suggests itself
[34], [35]. High-performance processors need two kinds of
wires. First, there are the short wires that serve the vast
majority of interconnects. For CMOS processors they are
typically up to 12 mm in length. They are mainly responsible for making the chip wirable by providing sufficient
number of interconnections. Here the
delay plays no
appreciable role; capacitance is the only consideration.
Such short wires should follow the minimum lithography
features of the available technology. Second, there is a
need for long wires, where density is secondary to delay
considerations. They run between distant parts of the chip,
and their characteristic length is that of a chip-edge. A good
scaling gauge for such long wires is that the time of
signal propagation on them should be only a small fraction
of the cycle-time. From such considerations, it immediately
follows that the cross section of these wires and insulators
cannot have minimum lithography features. This type of
interconnects will be referred to as fat wires [34], [35].
Typically, the bottom two levels are at the finest pitch for
which the device technology can take advantage. Here, lines
and spaces should be almost at minimum design rules. The
next two levels dimensions should already pay attention to
the
problem, and finally the top two can serve to run
signals to full chip-edge length, or longer, distances. With
this type of wiring, where conductor and dielectric crosssectional dimensions are scaled together, capacitance per
unit length stays constant for each level, while resistance
decreases proportionally with wire cross-sectional increase.
The number of wiring planes and pitch ratios amongst them
must be optimized for any given design. For reaching the
highest performances, ratios where the second - planepair resistance per unit length is one-fourth and the third is
one-thirty-sixth that of the bottom plane are quite realistic.
TAUR et al.: CMOS SCALING INTO THE NANOMETER REGIME
Fig. 19. RC delay versus wire length for three different wire sizes
(assuming square wire cross sections). Wires become electromagnetic-wave-propagation limited when the RC delay equals the time
of flight over the line length.
SOI MOSFETs exhibit superior subthreshold characteristics, thereby attracting attention for low-voltage operation.
However, the nearly ideal subthreshold slope of a FD
SOI MOSFET results from very different factors when
compared to a double-gate MOSFET, which also has nearly
ideal subthreshold slope [37]. The channel potential of any
MOSFET is controlled by both the front gate and the back
gate. For double-gate MOSFETs, the ideal subthreshold
slope is because the channel is being controlled by both
gates, which are switched simultaneously. However, in FD
SOI MOSFETs, the nearly ideal subthreshold slope occurs
because the back gate does not have much control over
the channel. The lack of back gate control leaves a wide
depletion or dielectric region vulnerable to source-drain
field penetration, which results in poor short-channel effect
[38]. This point is further addressed in Section IV-D.
It should be noted that there is no abrupt boundary
between PD and FD; the floating-body or kink effect occurs
PROCEEDINGS OF THE IEEE, VOL. 85, NO. 4, APRIL 1997
Fig. 24. Layer cross section of a Si/SiGe CMOS structure. The strained Si channel is populated
with electrons under positive gate bias (see conduction band on the right), and the strained SiGe
channel is populated with holes under negative gate bias (see the valence band on the left).
(a)
Fig. 25. Measured waveform of a 43-stage, 0.08-m channel
n-MOS ring oscillator at 85 K. The gate delay is 7.8 ps per stage.
(b)
Fig. 26. Simulated threshold voltage versus channel length, comparing short-channel effect of double-gate FETs (solid lines)
with SOI MOSFETs (dashed lines), where the threshold of the
long-channel FETs has been taken as zero. These values are extracted from drift diffusion simulations of the subthreshold regime
of these FETs. Inset: cross-sectional structure of a double-gated
FET.
a 50-A-thick
channel and 30-A-thick
gate oxide, Fig. 26
indicates a minimum channel length of 30 nm using the
criterion of 100-mV threshold variation for a 30% gate
length variation. Similarly, for an oxide thickness of 20
and a channel thickness of 40 A,
a minimum channel
A
length of 2025 nm should be possible. To avoid threshold
fluctuations due to the discreteness of the dopants, it would
be necessary to adjust the threshold of this FET by the work
function of the gate, leaving the channel undoped.
To evaluate the potential on-state performance of these
FETs, detailed Monte Carlo simulations were performed
[51], [52] using the simulator DAMOCLES [53]. Both nand p-channel MOSFETs have been simulated, yielding
499
Fig. 28. Monte Carlo simulation of electron energy versus position down the channel of an
n-channel double-gate MOSFET. The points represent electrons and the line indicates the conduction
band edge. The height of the points above the band edge indicates their kinetic energy.
low output conductance, high-performance IV characteristics for both device types, as is illustrated in Fig. 27. The
transconductance exceeds 2300 mS/mm for this nFET, and
it reaches 1300 mS/mm for the pFET. Transient Monte
Carlo simulations were also done for the nFET switching
a capacitive load equivalent to another nFET. This resulted
in a minimum estimated switching time of 1.1 ps for this
nFET, clearly indicating the potential for performance in
these tiny FETs.
The Monte Carlo simulations also allow an analysis of the
internal carrier behavior of the double-gate MOSFET. As
illustrated in Fig. 28, the carriers behave quite ballistically
in these short devices. Very little kinetic energy is lost until
500
(a)
(b)
(c)
Fig. 29. Three possible orientations of a double-gate MOSFET
on a silicon wafer. Examples of devices fabricated are: (a) Colinge
et al. [54], Tanaka et al. [55], (b) Takato et al. [56], and (c)
Hisamoto et al. [57].
In conclusion, CMOS devices will scale into the nanometer regime with improved device performance and lower
power before running into fundamental barriers of physics.
A 50-nm channel length bulk MOSFET would have a
20-A-thick
gate oxide and a power supply voltage near
1 V for high-performance applications. The maximum
electric field would reach 5 MV/cm in the oxide and 2
MV/cm in silicon ( 1 MV/cm at threshold condition).
Nonuniform channel doping in both the vertical (retrograde)
and the lateral (halo) directions will be used to minimize
short-channel effects. For low-power applications, multiple
and/or dynamic threshold voltages may be implemented
with a power supply voltage 1 V for active/standby power
management. A hierarchical wiring scheme with sixeight
metal levels will be necessary to deal with the interconnect
delays. With the top-level global wires limited only by
the speed of electromagnetic wave propagation, it should
be possible to build processors with clock frequencies well
into the gigahertz regime.
In practice, however, many difficult challenges lie ahead
in tightening process tolerances to satisfy more stringent
defect density and reliability requirements in future generation CMOS technologies. Some of the solutions call for a
paradigm shift and costly buildup of new infrastructures,
e.g., in the case of X-ray lithography. Others require
near atomic-level thickness control and nanometer-scale
lateral-dimension inspection and control. Much lower defect densities and higher device yields than todays standard
will undoubtedly be required when multi-billion transistors
are fabricated on a single chip.
Finally, several novel/alternative CMOS device structures have been discussed. SOI and SiGe devices offer
performance and power advantage over bulk CMOS without channel length scaling, while low-temperature CMOS
and double-gate MOSFETs can potentially take us to
the outermost limit of silicon scaling. The challenges,
however, lie in the fabrication of self-aligned double-gate
MOSFETs and low-cost cooling of VLSI chips/packages
to low temperatures.
ACKNOWLEDGMENT
The authors would like to thank R. H. Dennard, T. H.
Ning, L. Su, F. Assaderaghi, K. A. Jenkins, S. Rishton, J.
O. Chu, J. Y.-C. Sun, E. Crabbe, E. Nowak, and F. Stern
for many stimulating discussions. They would also like to
thank the Yorktown, NY, Silicon Facility and the Advanced
Silicon Technology Center in East Fishkill, NY, for device
fabrication.
501
REFERENCES
[1] Y. Taur, S. Wind, Y. Mii, Y. Lii, D. Moy, K. Jenkins, C. L.
Chen, P. J. Coane, D. Klaus, J. Bucchignano, M. Rosenfield, M.
Thomson, and M. Polcari, High performance 0.1 m CMOS
devices with 1.5 V power supply, in 1993 IEDM Tech. Dig.,
pp. 127130.
[2] Y. Taur, D. S. Zicherman, D. R. Lombardi, P. J. Restle, C.
H. Hsu, H. I. Hanafi, M. R. Wordeman, B. Davari, and G. G.
Shahidi, A new shift and ratio method for MOSFET channel
length extraction, IEEE Electron Device Lett., vol. 13, p. 267,
1992.
[3] S. C. Palmateer, R. R. Kunz, M. W. Horn, A. R. Forte, and
M. Rothschild, Optimization of a 193-nm silylation process
for sub-0.25 m lithography, in Proc. SPIE, 1995, vol. 2438,
p. 455.
[4] P. Agnello, T. Newman, E. Crabbe, S. Subbanna, E. Ganin,
L. Liebmann, J. Comfort, and D. Sunderland, Phase edge
lithography for sub-0.1 m electrical channel length in a
200 mm full CMOS process, IEEE 1995 Symp. on VLSI
Technology, Dig. Tech. Papers, 1995, p. 79.
[5] Y. Chen, R. K. Kupke, F. Rousseaux, F. Carcenac, D. Decanini,
M. F. Ravet, and H. Launois, 50 nm X-ray lithography using
synchrotron radiation, J. Vac. Sci. Technol., vol. B12, p. 3959,
1994.
[6] H. I. Smith and M. L. Schattenburg, X-ray lithography, from
500 to 30 nm: X-ray nanolithography, IBM J. Res. Devel., vol.
37, p. 319, 1993.
[7] R. H. Stulen, 13-nm extreme ultraviolet lithography, IEEE J.
Select. Topics. Quantum Electron., vol. 1, p. 970, 1995.
[8] W. Langheinrich, A. Vescan, B. Spangenberg, and H. Beneking,
The resolution of the inorganic electron beam resist, LiF(Al3 ),
Microelectronics Eng., vol. 23, p. 287, 1994.
[9] M. Parikh, Corrections to proximity effects in electron beam
lithography, J. Appl. Phys., vol. 50, p. 4371, 1979.
[10] S. J. Wind, C. M. Reeves, J. J. Bucchignano, Y. T. Lii, T.
H. Newman, D. P. Klaus, J. Keller, R. P. Volant, B. Tebin,
and F. J. Hohn, Fabrication of compact 100 nm-scale silicon
metal-oxide-semiconductor field effect transistrs, J. Vac. Sci.
Technol., vol. B10, p. 2912, 1992.
[11] H. C. Pfeiffer and W. Stickel, PREVAILAn e-beam stepper
with variable axis immersion lenses, Microelectronic Eng., vol.
27, p. 143, 1995.
[12] S. D. Berger, J. M. Gibson, R. M. Camarda, R. C. Farrow, H.
A. Huggins, J. S. Kraus, and J. A. Liddle, Projection electron
beam lithography: A new approach, J. Vac. Sci. Technol., vol.
B9, p. 2996, 1991.
[13] L. R. Harriott, S. D. Berger, J. A. Liddle, G. P. Watson, and
M. M. Mkrtchyan, Space charge effects in projection charged
particle lithography systems, J. Vac. Sci. Technol., vol. B13,
p. 2404, 1995.
[14] S. C. Minne, H. T. Soh, P. Flueckiger, and C. F. Quate,
Fabrication of 0.1 m metal oxide semiconductor field effect
transistors with the atomic force microscope, Appl. Phys. Lett.,
vol. 66, p. 703, 1995.
[15] R. H. Dennard, F. H. Gaensslen, H. N. Yu, V. L. Rideout,
E. Bassous, and A. R. LeBlanc, Design of ion-implanted
MOSFETs with very small physical dimensions, IEEE J.
Solid-State Circuits, vol. SC-9, p. 256, 1974.
[16] Y. Mii, S. Wind, Y. Taur, Y. Lii, D. Klaus, and J. Bucchignano,
An ultra-low power 0.1 m CMOS, in 1994 VLSI Technology
Symp. Tech. Dig., pp. 910.
[17] Y. Taur, Y. Mii, D. Frank, H.-S. Wong, D. Buchanan, S. Wind,
S. Rishton, G. Sai-Halasz, and E. Nowak, CMOS scaling into
the 21st century: 0.1 m and beyond, IBM J. Res. Devel., vol.
39, p. 245, 1995.
[18] Y. Mii, S. Rishton, Y. Taur, D. Kern, T. Lii, K. Lee, K.
Jenkins, D. Quinlan, T. Brown Jr., D. Danner, F. Sewell,
and M. Polcari, Experimental high performance sub-0.1 m
channel nMOSFETs, IEEE Electron Device Lett., vol. 15, p.
28, 1994.
[19] L. Su et al., A High performance 0.08 m CMOS, in 1996
VLSI Technology Symp. Tech. Dig., p. 12.
[20] F. Stern, Self-consistent results for -type Si inversion layers,
Phys. Rev. B, vol. 5, no. 12, pp. 48914899, 1972.
[21] D. A. Buchanan and S.-H. Lo, Growth, characterization and
the limits of ultra-thin SiO2 -based dielectrics for future CMOS
applications, in The Physics and Chemistry of SiO2 and the
Si-SiO2 Interface-3, H. Z. Massoud, E. H. Poindexter, and C.
502
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
Douglas A. Buchanan (Member, IEEE) received the B.Sc. and M.Sc. degrees in electrical
engineering from the University of Manitoba,
Winnipeg, Canada, in 1981 and 1982, respectively. In 1986, he received the Ph.D. degree
from Durham University, Durham, U.K.
Following a two-year Post-Doctoral Fellowship at IBMs Thomas J. Watson Reaearch Center, Yorktown Heights, NY, he spent three years
in the CVD thin-film technology group in the
IBM Microelectronics Division. He currently
works on issues pertaining to grown and characterization of ultra-thin
dielectrics at the Thomas J. Watson Research Center.
George A. Sai-Halasz (Fellow, IEEE) graduated in physics from the Eotvos Lorand University, Budapest, Hungary, in 1966. He received
the Ph.D. degree in physics from Case Western
Reserve University, Cleveland, OH, in 1972.
Over the next two years, he held a PostDoctoral Fellowship at the Physics Department
of the University of Pennsylvania, Philadelphia.
He joined the IBM Thomas J. Watson Research
Center in 1974. He has made seminal experimental/theoretical contributions in field ranging
from basic science to technology. He contributed to quantum solids,
nonequilibrium superconductivity, and the physics and device aspects
of semiconductor lattices. He was one of the originators and primary
theoretical exponent of the Type II superlattice system. He invented and
developed a statistical modeling scheme for predicting radiation-induced
soft-error rates in VLSI circuits. He was the manager and technical
leader of the first successful effort to demonstrate 1-m and below FET
technology feasibility. His current interests are in the areas of systems
and high-end CMOS processor design.
Dr. Sai-Halasz was co-recipient of the 1997 IEEE Cledo Brunetti Award.
504