Vous êtes sur la page 1sur 18

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 57, NO.

10, OCTOBER 2010 2363

Physics, Technology, and Modeling of


Complementary Asymmetric MOSFETs
Constantin Bulucea, Fellow, IEEE, Sandeep R. Bahl, Senior Member, IEEE, William D. French,
Jeng-Jiun Yang, Member, IEEE, Pascale Francis, Tikno Harjono, Vijay Krishnamurthy,
Jon Tao, and Courtney Parker, Member, IEEE

Abstract—The physics, technology, and modeling of comple- I. I NTRODUCTION


mentary asymmetric MOSFETs are reviewed and illustrated with
statistically representative silicon data from a recent manufactur-
ing implementation, in which the transistors for the secondary
power supply voltage are offered in asymmetric and symmet-
H ISTORICALLY, the planar MOSFET has been the first
major active device to be built and used in symmet-
ric input–output construction. This has been possible due its
ric constructions. The in-depth analysis of the device physics of input–output reversibility, allowing the simplest implementa-
asymmetric transistors provides new insights into their physical
operation and into the operation of transistors using halo im- tion of bidirectional sampling switches, and due also to its un-
plants in general. The variability, matching, and noise implica- precedented functional density in large-scale-integration (LSI)
tions of using halo implants are also analyzed, concluding that circuits.
both asymmetric and symmetric devices need to be offered for The circuit and device structure requirements for the in-
uncompromised circuit design. The challenges associated with the put and output terminals of amplifying or switching devices
compact modeling the asymmetric transistors are also reviewed
and illustrated. The preferred manufacturing implementation are different in most applications, particularly in analog and
uses retrograde wells with no dopant fillers at the surface, while mixed-signal applications. This means that using symmetric
avoiding the drain-to-source punch-through by source-side-only MOSFETs implies design compromises and diminished perfor-
halo implants. In addition to the known switching speed and mance in comparison with the ideal situation of using optimized
maximum voltage gain advantages of the asymmetric transis- asymmetric devices. Still, the economics of large-scale integra-
tors, this particular device architecture offers superior hot-carrier
reliability and transistor design flexibility. The availability of ret- tion has dictated the use of symmetric MOSFETs.
rograde wells enables construction of high-reliability complemen- However, with advanced scaling, the output characteristics of
tary extended-drain MOSFETs for a third higher power supply the transistor degrade toward diminished voltage gain, which is
voltage. critical to analog design. In large-scale integration of digital and
Index Terms—Analog, antipunch-through, asymmetric, analog/mixed-signal/RF circuits such as in system-on-a-chip
BSIM3, compact modeling, CMOS, cutoff frequency, deep and systems, the economic balance changes in favor of accepting
lightly doped drain extension (DLDD), doping-gradient-induced the additional costs (masks and implants) for building better
barrier lowering (DGIBL), doping profile, drain-extended MOS, optimized devices, such as asymmetric MOSFETs [1]–[4].
drain-induced barrier lowering (DIBL), drift-diffusion, dual gate
oxide (DGO), empty-well, extended-drain, extended-drain MOS The asymmetric MOSFETs are not new. The MOS laterally
(XD-MOS), gate-induced drain leakage, gate-induced source diffused transistors of both planar [5], [6] [laterally diffused
leakage (GISL), graded-channel, halo, high-voltage, hot-carrier, MOS (LDMOS), vertical diffused MOS (VDMOS)] and trench
large-scale-integration (LSI), laterally diffused MOS (LDMOS), (trench VDMOS) [7]–[9] construction are examples of well
matching, Medici, MOSFET, mixed-signal, noise, output known asymmetric power MOSFETs where the channels are
resistance, reflection coefficient, reliability, retrograde well,
scattering theory, single-pocket, STI, subcircuit, switching speed, laterally nonuniform (diffused), and the source and drain con-
system-on-a-chip, T-gate, threshold adjust, transconductance, structions are different.
trench, TSuprem-4, variability, vertical diffused MOS (VDMOS), Asymmetric low-power MOSFETs using various implemen-
voltage gain. tations of graded channels or different source and drain re-
gions, or combinations of them, have also been proposed, built,
and analyzed [10]–[41]. In general, discussions emphasize the
advantages of asymmetric transistors over their conventional
symmetric counterparts, with little or no analysis of their lim-
itations and the optimum use in actual product development.
Manuscript received March 22, 2010; revised June 18, 2010; accepted
June 21, 2010. Date of publication August 5, 2010; date of current version The terminology used to describe this class of MOSFETs is
September 22, 2010. The initial phase of this work was supported in part by quite nonuniform and includes designations such as asymmetric
the State of California through matched funding under the Microelectronics (or asymmetrical), asymmetric halo, graded channel, lateral
Innovation and Computer Research Opportunities Research Project 04-090.
The review of this paper was arranged by Editor A. Schenk. asymmetric channel, laterally doped channel, single halo, sin-
The authors are with National Semiconductor Corporation, Santa Clara, gle pocket, etc. Although each designation has its own merits in
CA 95052-8090 USA (e-mail: constantin.bulucea@nsc.com). describing the nature of these transistors, we will use the simple
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org. asymmetric MOSFET designation. We restrict it not to include
Digital Object Identifier 10.1109/TED.2010.2057197 the LDMOS and VDMOS power transistors. Also, we restrict

0018-9383/$26.00 © 2010 IEEE


2364 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 57, NO. 10, OCTOBER 2010

our review and analysis to bulk devices, thus excluding silicon- II. D EVICE AND P ROCESS
on-insulator implementations [42]–[46]. A RCHITECTURE C ONSIDERATIONS
A conventional symmetric long-channel transistor is shut off
An asymmetric MOSFET can be generally described as a
over the entire length of its uniformly doped channel at gate
planar MOSFET having the following: 1) a laterally nonuni-
biases below the threshold voltage. At a closer analysis, the
form body profile (channel asymmetry) with or without
current flow in the channel only needs to be interrupted at
2) different source and drain regions (source/drain asymmetry).
a single location between the source and drain for the drain
Laterally nonuniform body profiles are typically implemented
current to be turned off. This can be done by providing the
with halo implants on the source-side of the channel, while
necessary threshold doping at that location.
different source and drain regions are created with different
In a systematic 2-D computer simulation analysis of the
masked implants over the source and drain regions or their
laterally nonuniformly doped long-channel MOSFET,
extensions.
Stockinger et al. [24] varied the position of the maximum
Attempting to build MOSFETs with channel asymmetry at
channel doping between the source and drain (i.e., the position
the minimum printable gate length typically involves aligning
where the channel current is interrupted), demonstrating that
of one mask edge to a position around the middle of the
the best drive/leakage performance is obtained when the
polygate regions, which is already printed at the limits of the
maximum channel doping is near the source and calling that
lithography capability. Although this has been the most typical
transistor structure “the peak device.”
implementation in research/development efforts, applying it to
The situation is more subtle in short-channel devices where
manufacturing is not a trivial task.
the graded nature of the channel doping differentiates the
A more realistic approach to exploiting the advantages of
asymmetric transistors from their symmetric counterparts.
asymmetric MOSFETs is to implement them for the secondary
Lundstrom’s scattering theory of the short-channel MOSFET
bias voltage of a given CMOS process, which is typically
[47]–[49] can be used in the analysis, as later discussed.
referred to as dual voltage or dual gate oxide (DGO) CMOS
Several circuit applications have been reported [50]–[63].
[64], [65]. The secondary voltage capability is obtained with
Yet, many subtleties of internal operation, compact modeling,
a longer gate and a thicker gate oxide than those of the pri-
and optimal use have largely remained uncovered, which has
mary bias voltage. For example, in a scaled CMOS process
precluded their entering into the mainstream circuit design.
designed for a 1.2-V primary power supply voltage, having
This paper reviews the physics, technology, and modeling of
0.13-μm minimum gate length and ∼2-nm physical gate oxide
complementary asymmetric MOSFETs specifically integrated
thickness, the asymmetric MOSFETs would be implemented
for analog and mixed-signal applications, with new insights
with gate lengths of 0.25 μm or more and a gate oxide of
from a recent development of a manufacturing process.
∼6 nm, allowing a secondary bias voltage of 2.5 V or more.
A brief discussion of the available choices in devising an
Obviously, aligning the mask edges of the halo or source/
asymmetric MOSFET architecture in terms of device, process,
drain extension implants to positions around the middle of the
and power supply voltages in Section II leads to our preferred
gate is no longer a lithography challenge at gate lengths that are
empty well implementation, used for illustration throughout the
at least two times larger than the minimum printable values.
entire paper.
Another major architecture consideration is the choice of the
Section III presents a top-level review of the advantages and
type of the doping profile of the well in which transistors are
challenges of asymmetric MOSFETs, in comparison with their
built. Typically, the well profiles of symmetric DGO transistors
symmetric counterparts, intentionally keeping the discussion at
in a scaled CMOS process are made of a high-energy main well
a qualitative level.
that is subsequently filled with lower energy implants designed
The manufacturing of the asymmetric and symmetric
to fend off the channel punch-through (antipunch-through im-
MOSFETs in the aforementioned preferred architecture is out-
plant) and to insure the nominal threshold voltage of the process
lined in Section IV, followed by a discussion of the limits
(threshold adjust implant). For CMOS processes with gate
in boosting the drive capability of the asymmetric MOSFETs
lengths larger than 0.25 μm, the halo implants are not required,
in Section V. Section VI analyzes the device physics of the
so that the threshold voltage VT is roughly determined as
asymmetric transistors in greater depth, using a combination
of drift-diffusion TCAD and physical reasoning based on the VT = αQV T Adjust + βQAP T (1)
scattering theory of the short-channel MOSFETs.
Section VII is dedicated to the analysis, based on statistically where QV T Adjust and QAP T are the areal doses of the
validated silicon data, of the switching speed performance and threshold adjust and antipunch-through implants, respectively,
scaling, followed by a similar analysis of the small-signal and α and β are locally calibrated geometry- and energy-
speed in Section VIII. The variability, matching, and noise dependent coefficients measured in voltage/dose units. The
are discussed in Section IX. All these analyses are carried out contribution of the main well implant has been neglected in (1)
comparatively for asymmetric and symmetric MOSFETs. due to the typical retrograde nature of this implant in modern
The challenges of and solutions to the compact modeling of CMOS processes, which implies that little or no dopant from
the asymmetric transistors are briefly discussed in Section X. this well reaches within the depth of the depletion region under
Section XI describes the implementation of high-voltage the channel.
extended-drain transistors in our preferred architecture, illus- The simplest and most widely used way to convert a nonhalo
trating their cost and reliability advantages. symmetric MOSFET to an asymmetric, or graded-channel,
BULUCEA et al.: PHYSICS, TECHNOLOGY, AND MODELING OF COMPLEMENTARY ASYMMETRIC MOSFETs 2365

Fig. 2. First-order compact model of an asymmetric MOSFET. Q1 and Q0


are the threshold and main-channel transistor, of the two-transistor subcircuit,
Fig. 1. Cross-section of an asymmetric n-channel MOSFET using the empty-
respectively.
well architecture. The p-channel counterpart is complementary to the one
shown.
series and the gates and body regions each tied together, as
counterpart is to add a halo implant at the source end of the represented in Fig. 2.
channel. Obviously, this will add a third term, γQHalo , to The schematic representation of the device structure in Fig. 1,
(1), thus increasing the threshold voltage. Hence, the threshold along with the compact model in Fig. 2, is sufficient for a
adjust or the antipunch-through dose, or both, needs to be first-order physical discussion of the essential features of the
reduced to insure the same threshold voltage. The antipunch- asymmetric MOSFETs in comparison with their symmetric
through and threshold adjust implants can be both eliminated, counterparts. The symmetric MOSFETs used in our compari-
thus leaving the relatively stronger halo pocket at the source, son later are nonhalo transistors built in the same process (see
with the threshold voltage determined as Section IV), having the same gate oxide thickness, channel
width and channel length, and similar threshold voltages. The
VT = γQHalo . (2) discussion is intentionally kept at a qualitative level.

The aforementioned process architecture has been preferred A. Greatly Improved Hot-Carrier Reliability
in the process development used for illustration in this paper,
with the observation of a substantial increase of the channel With no halo at the drain side of the channel and with a
mobilities through diminished carrier-impurity scattering (next deeper lightly doped extension, the electric field in the drain
section). This architecture will be heretofore referred to as region, in normal operation, is lower than in a symmetric tran-
empty-well architecture for brevity.1 sistor. Therefore, at high drain voltages, the transistor operates
farther away from the avalanche breakdown condition at the
drain-well junction, hence with fewer hot carriers around the
III. S TRUCTURE AND F IRST-O RDER P HYSICAL surface.
M ODELING OF A SYMMETRIC MOSFETs The aforementioned improvement is substantial. For a
The basic device structure of an n-channel MOSFET with quantitative illustration, our typical 3-V/0.3-μm asymmetric
channel and source–drain asymmetry using the empty-well n-channel transistors built in empty-well architecture measure
architecture is described in Fig. 1. around 400 years extrapolated hot-carrier lifetimes for either
The channel asymmetry is created by using a halo implant VT or IDsat (drain saturation current) degradation. These
only on the source-side, while the source–drain asymmetry is lifetimes compare to 25 and 10 years in typical 3-V/0.4-μm
created by using a deep and lightly doped drain extension (deep symmetric transistors, for VT and IDsat , respectively, at the
LDD, or DLDD) and a conventional or more heavily doped same threshold voltage and voltage stressing conditions. The
source extension. Optionally, the n-channel transistors can be large reliability margin of the asymmetric transistors points to
isolated from the p-type substrate with a deep n well. The the possibility to use them at subminimum DGO gate lengths,
p-channel structure is perfectly complementary to its n-channel e.g., 0.25 instead of 0.3 μm, for improved speed performance,
counterpart except that it does not have, nor does it need, a deep as discussed later in Section VII.
well isolation.
In a first-order compact model representation, this transistor B. Increased Drain Saturation (“Drive”) Current
can be seen as a combination of a threshold transistor, Q1 ,
In long-channel devices, at a given gate voltage overdrive, the
representing the portion of the channel containing the graded
main-channel transistor of the two-transistor model, Q0 , having
channel implant, and a main-channel transistor, Q0 , represent-
a weakly doped channel, is in a stronger inversion condition
ing the rest of the channel, with the channels connected in
than the threshold transistor Q1 . Therefore, it contributes with a
smaller channel resistance to the total resistance of the channel
1 In sub-0.25-μm symmetric MOSFETs using halo implants, the threshold
in comparison with the situation when the entire channel is
adjust implant can be safely dropped [66], which actually improves the switch-
ing speed. The antipunch-through implant can be dropped too, but severe long- doped at the level required for setting the threshold voltage.
range reverse-short-channel effects make this option less attractive. Compared at the same gate voltage overdrive and drain voltage,
2366 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 57, NO. 10, OCTOBER 2010

the drain current of the asymmetric transistor is substantially


larger than the one in its symmetric counterpart.
In short-channel devices, the scattering theory of the short-
channel MOSFET [47]–[49], calling for an increased electric
field within a mean free path from the source-channel barrier
can be used to understand the observed increased drive current
[38]. In the absence of the drain halo, the great majority of
the field lines originating from the positively ionized donors
in the drain region terminate on the oppositely charged ionized
acceptor in the source halo, increasing the electric field there
(n-channel case).
In our classification of transistors into long- and short-
channel, a short-channel transistor is conventionally defined
around the minimum gate length that can be controlled in
lithography and etching, or used with acceptable hot-carrier
reliability, in the manufacturing of the respective transistors. In
our review illustration, the minimum gate length is 0.13 μm
(lithography control limit) for the first voltage (1.2 V) and
0.30 μm (conservatively defined reliability limit) for the sec-
ondary voltage (3 V) of the CMOS process. Consequently,
secondary-voltage transistors with gate lengths around 0.30 μm
are referred to as short-channel transistors.

C. Larger Drain Saturation Voltage in Long-Channel Devices


A larger drain saturation voltage is observed in long-channel
MOSFETs, which is an undesirable property. This is a direct
consequence of the weak body doping around the drain region
caused by the absence of a halo implant there. As the drain
voltage is increased, the drain field easily pushes the boundary
of the drain depletion region at the surface, toward the source, Fig. 3. (a) Typical output I–V characteristics of (thick lines) long-channel
reducing the effective length of the channel and delaying the on- asymmetric and symmetric n-channel MOSFETs with L = 5 μm and W =
set of the pinch-off condition. This trend eventually stops when 10 μm. (b) Typical output I–V characteristics of short-channel asymmetric and
symmetric MOSFETs with L = 0.3 μm and W = 10 μm. The arrows indicate
the boundary of the heavily doped source halo is approached, the drain current increase in asymmetric transistors compared to symmetric
and the device enters a region of strong saturation. counterparts biased at same gate overdrive, as marked. All characteristics are
In short-channel asymmetric transistors, this effect is less at zero body bias, VB = 0 V.
visible and may even be reversed.
the exact definition of VDsat , it is obvious from this illustra-
tion that in long-channel asymmetric MOSFETs of same gate
D. Improved Saturation of the Drain Current (Larger Output length, L and width W , VDsat is larger than the one calculated
Resistance in Saturation) with the commonly used approximation (3).
Once the transistor enters the drain saturation regime, the
drain field can no longer push into the source halo region, E. Increased Linear and Saturation Transconductance
and a stronger saturation sets in (larger output resistance) in
comparison with a symmetric transistor. Moreover, with the Both long- and short-channel transistors turn on more effi-
reduced field at the drain, the drain current has little or no ciently in asymmetric construction, offering an increased linear
transconductance, gm , with a higher peak value at the inflection
tendency to creep up due to impact ionization toward the end
of the drain voltage range. point of the ID (VD ) characteristics, as shown in Fig. 4(a)
(long-channel) and Fig. 4(b) (short-channel).
Properties B–D are illustrated in Fig. 3(a) (long-channel) and
In short-channel devices built at the minimum gate length,
Fig. 3(b) (short-channel), based on our typical 3 V silicon. The
long-channel threshold voltages are 0.62 and 0.53 V for the the transconductance is flatter (less peaked) around the maxi-
mum value, hence easier to capitalize on in applications (see
asymmetric and symmetric transistors, respectively.
Section X).
The drain saturation voltage shown in these figures is con-
ventionally defined as The increased linear transconductance property is more sub-
tle in nature. No first-order model is proposed to explain it other
VDsat ≈ VG − VT (3) than observing that, for monotonically increasing ID (VG , VD )
functions, this property is mathematically expected, except
for symmetric transistors [67], [68] and is calculated with an for the unlikely situation when the drain current increase in
empirical equation for asymmetric counterparts. Regardless of asymmetric devices is the same at all drain voltages.
BULUCEA et al.: PHYSICS, TECHNOLOGY, AND MODELING OF COMPLEMENTARY ASYMMETRIC MOSFETs 2367

is expected to be higher in transistors having channel asym-


metry than in their symmetric counterparts. With channel pro-
file changes leading to reduced gate capacitances, the cutoff
frequency is even higher. The maximum oscillation frequency
fmax being related to fT [68], tends to follow the same trend.
A more detailed discussion of these properties is provided in
Section VIII.

H. Larger Parameter Variability and Noise


The asymmetric transistors have larger parameter variability
and noise in comparison with their symmetric counterparts that
do not use halo implants. These limitations are discussed in
Section IX.

IV. M ANUFACTURING OF A SYMMETRIC AND S YMMETRIC


MOSFETs IN E MPTY-W ELL P ROCESS A RCHITECTURE
The fabrication process of the asymmetric MOSFETs used
in the introductory discussion before consists of a conventional
submicrometer DGO CMOS process with changes to provide
the following: 1) retrograde substrate wells (empty-well archi-
tecture); 2) graded channel; and 3) deeper and more lightly
doped drain extensions (DLDD) than the source extensions
for the DGO transistors, as represented in Fig. 5. Unlike a
conventional halo implant, the graded-channel (halo) implant
is applied only on the source-side of the channel, through a
dedicated mask.
The minimum gate length of this process is 0.13 μm, which
corresponds to a primary supply voltage and electrical gate ox-
ide thickness in inversion of 1.2 V and 2.8 nm, respectively. The
Fig. 4. (a) Typical transfer I–V characteristics of (thick lines) long-channel
asymmetric and symmetric n-channel MOSFETs with L = 5 μm and W = secondary supply voltage and electrical gate oxide thickness in
10 μm. (b) Typical transfer I–V characteristics of short-channel asymmetric inversion are 3.0 V and 6.2 nm, respectively. Both oxides are
and symmetric MOSFETs with L = 0.3 μm and W = 10 μm. In all cases, the grown using an in situ steam generation process [69].
drain current of the asymmetric transistors increases more rapidly around the
inflection point of the transfer characteristic, providing substantial transconduc- Aside from allowing the manufacturing of high-performance
tance advantage. All characteristics are at VB = 0 V. reliable asymmetric transistors, the empty-well architecture has
the advantage of enabling multiple combinations of well and
The asymmetric transistors also show a higher saturation LDD/halo implants for manufacturing of several types of “free
transconductance, gmsat . devices,” such as symmetric DGO MOSFETs, low-threshold-
voltage MOSFETs, low-cost bipolar transistors, complemen-
tary high-voltage MOSFETs, etc.
F. Increased Maximum Voltage Gain The complementary symmetric MOSFETs are particularly
Combination of increased output resistance (property D) and required in the design of the bidirectional sampling switches
increased saturation transconductance (property E) result in an (transmission gates or T-gates), where the drain and source
increased maximum voltage gain functions change during circuit operation as the voltage applied
at the input of the switch increases or decreases.
AMax = gmsat · rout (4) The simplest approach to integrate symmetric transistors into
a process architecture that is primarily designed for asymmetric
which is particularly important in analog design. transistors is to apply the graded channel implant at both the
source and drain ends of the channel. Although used in the
past [54], this approach tends to yield unreliable symmetric
G. Improved Small-Signal High-Frequency Behavior transistors due to the high electric field in the drain region. The
Deriving directly from the transconductance comparison and doping concentration of the graded channel implant is higher
assuming similar gate capacitances, Cg , the small-signal cutoff than that of a conventional halo implant, which intensifies the
frequency [67] electric field around the drain end of the channel.
Differently from the aforementioned transistors, reliable
gm symmetric transistors are formed if the well profiles of the low-
fT ≈ (5)
Cg voltage transistors, filled with all their implant components,
2368 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 57, NO. 10, OCTOBER 2010

Fig. 5. Simplified representation of the process architecture used in manufacturing transistors with channel and source–drain asymmetry in empty-well
architecture.

characteristics of these symmetric transistors are similar to


those of conventional foundry DGO transistors at comparable
voltage specifications [70].
Extended-drain transistors capable of operating at 12 V drain
voltage can be also manufactured as “free devices” through
combination of existing implants and masks, as separately
discussed in Section XI.

V. H OW M UCH B ETTER D RIVERS THE A SYMMETRIC


T RANSISTORS C AN B E IN C OMPARISON W ITH
T HEIR S YMMETRIC C OUNTERPARTS ?
As shown in Section III, the asymmetric transistors are better
Fig. 6. Cross-section of a symmetric n-channel MOSFET combining the
filled wells of the low-voltage transistors and the deep lightly doped drain
drivers, at given gate overdrive, than their symmetric counter-
extensions of the asymmetric empty-well transistors. The p-channel MOSFET parts. The asymmetry can be further enhanced by providing
is perfectly complementary to the one shown. more heavily doped source extensions, thus reducing the para-
sitic voltage drop on the source series resistance. This results in
are combined with the deep LDD implants of the asymmetric additional increase of the drive current, which naturally leads
transistors, as shown in Fig. 6. Compatible threshold voltages to the question on how far the asymmetric drive performance
are obtained, without the use of halo implants, due to the can be pushed. The answer to this question depends on the
relatively higher doping levels in the low-voltage wells. The specifics of the applications in which these transistors are used,
minimum gate lengths of transistors so built may be different as described below.
from those of asymmetric counterparts due to manufacturing In applications where the body terminal is permanently con-
and hot-carrier reliability requirements. For example, in our nected to the source terminal and the drain is always at a higher
preferred empty-well architecture, the asymmetric transistors potential than the source, the source extension can be doped as
had a minimum gate length of 0.3 μm for convenient manu- heavily as possible, with the benefit of greater drive current.
facturing, while symmetric were restricted to a minimum gate In applications where the body terminal is not connected to
length of 0.4 μm for hot-carrier-lifetime considerations. The any of the source, or the drain terminals and the drain terminal
BULUCEA et al.: PHYSICS, TECHNOLOGY, AND MODELING OF COMPLEMENTARY ASYMMETRIC MOSFETs 2369

Fig. 7. GISL in reverse drain–source bias as a function of the drain saturation


current in forward drain–source bias for a typical n-channel asymmetric
MOSFET with various conditions of the source extension implant.

Fig. 8. Three-dimensional representation of the doping profile. The vector


can be at a lower potential than the source, the source extension data based on the highly nonuniform mesh of the simulator was processed
needs to be doped at a moderate level to avoid the gate-induced into a regular matrix format, using bicubic spline interpolation, before the final
visualization processing in MathCAD.
drain junction leakage of the reverse transistor. Except for
inventive circuit designs exploiting specific characteristics of VI. I N -D EPTH D EVICE P HYSICS A NALYSIS
reversely biased transistors [62], [63], the intentional use of
such devices is not recommended due to their known degraded The typical I–V characteristics illustrated in Section III
punch-through behavior and hot-carrier reliability [11], [12]. show that the asymmetric transistors have larger drive cur-
However, transient situations of reverse biasing may occur rents and transconductance in comparison with their symmetric
in circuit operation. In such situations, the source and drain counterparts. In the following two sections, this property is
change their function in device operation, the source working discussed in greater physical detail.
as a drain and vice-versa.
Holding the source and drain designations of the transistor
A. Long-Channel Case
terminals unchanged, as determined by their layout, irrespec-
tive of relative bias, this effect will hereafter be referred to In long-channel devices, the increased drive current can be
as gate-induced source leakage, or GISL, for similarity with easily understood using the two-transistor model, as already
the gate-induced drain leakage [68]. For a custom-designed discussed in Section III. Moreover, the empty-channel nature
process, a tradeoff has to be made between having superior of the channel transistor Q0 (Fig. 2) suggests that, with fewer
driving capability in normal transistor operation and manage- impurities in the channel, the increased carrier mobility is
able source junction leakage in reverse transistor operation, expected to result in an increase in the drain current, at constant
when the source is biased at a higher potential than the drain. gate overdrive. In addition, an important portion of the drain
The GISL effect is best characterized through special drive- current increase is explained by the favorable field structure in
leakage characteristics where the source leakage with reverse these transistors from the point of carrier transmission over the
drain–source bias, is plotted, in logarithmic scale, against the source-channel barrier of the threshold transistor Q1 . The high
drain saturation current in forward drain–source bias, in linear field at the source end of the channel, required by the scattering
scale. Fig. 7 shows the GISL characteristics of our empty-well theory for maximizing the drain current [47]–[49], is present in
process for various implant conditions of the source extension these transistors.
implant. Being aware of the difficulties associated with probing the
The GISL current is measured with the source and drain tied limitations of the drift-diffusion theory in silicon, where the
together to differentiate the junction from the channel leakage, saturation velocity vsat , entering the drift-diffusion equations,
and the body terminal is grounded. The leakage through the and the thermal velocity vT , entering the scattering theory
drain junction has been determined to be unimportant from equations, happen to be similar, we will not attempt to prove
prior measurements on the same circuit with the drain con- the drift-diffusion theory wrong. On the contrary, we confirm
nected to the power supply voltage (normal drain biasing). that, with appropriate coefficient calibration, accurate matching
The raw data set has been reduced to a same-threshold-voltage between experimental data and the drift-diffusion simulations
equivalent data set through statistical modeling using the JMP can be obtained. Yet, the electrical field structure inside the
software [71]. One can see that the dose of the source ex- device, revealed by these simulations, provides considerable
tension implant is the primary controller of the drive/source- arguments in support of the scattering theory.
leakage tradeoff, while the energy and the angle play secondary Fig. 8 shows, in a 3-D representation, the net doping profile
roles. of a 1-μm asymmetric transistor resulted from the simulation of
2370 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 57, NO. 10, OCTOBER 2010

Fig. 9. Distribution of the electric potential and lateral electric field at the surface of an (a) symmetric and (b) asymmetric n-channel MOSFET built in the
empty-well architecture described in Section IV. Shaded areas indicate depletion regions. Notice the strong electron extracting field at the inside edge of the
source-channel depletion region of the asymmetric MOSFET.

our empty-well process using the TSuprem IV process simulator region, as expected in the absence of a halo doping there. A
[72] with coefficients calibrated to the actual profiles described strong electron extracting field is present at the channel edge
by corroborated spreading-resistance probe and secondary ion of the source-channel depletion region, which reduces back-
mass spectroscopy characterization. scattering of the electrons that have surmounted the potential
Fig. 9(a) and (b) represent the distribution of the electric barrier.
potential and longitudinal electric field at the surface, calcu- In view of the scattering theory, the source/source-
lated with the drift-diffusion equations using the Medici device extension/source-halo region can be seen as a near-ballistic
simulator [73] for VG = VT + 0.2 V and VD = 1.5 V, along transistor corresponding to the threshold transistor Q1 of the
with simplified 1-D representations of the device structure two-transistor compact model in Fig. 2. The function of a local
at the surface for the symmetric and asymmetric transistors, drain for this transistor is provided by the built-in field of
respectively. The source-channel portion of each device is the graded halo doping, which extracts the electrons over the
subdivided into Source (S), Source Extension (SE), and Source source-channel barrier. From there on, the extracted electrons
Halo (SH) regions of substantially different doping, in order are taken over by the conventional long-channel field-effect
to ease understanding of the field structure in terms of the mechanism of the main channel transistor Q0 in the two-
classical transistor theory. Similarly, the drain-channel portions transistor compact model.
are divided into Drain (D) and Drain Extension (DE) regions. In the aforementioned physical interpretation, it is important
At low current levels, the drift-diffusion simulations are to realize that the drive current of the asymmetric transistor is
believed to be fairly accurate, regardless of the drift-diffusion- limited by the carrier emission over the barrier and the scatter-
versus-scattering discussion, considering that the field is elec- ing phenomena taking place in transistor Q1 . In particular, the
trostatically determined by the doping structure and the applied magnitude of the built-in field in the source halo region, desir-
voltages. ably as high as possible, controls the channel backscattering (or
Starting with the simpler field situation of the symmetric reflection) coefficient r and ultimately the drive current of the
transistor [Fig. 9(a)], one can identify the essential features whole transistor.
of the field distribution. The two peaks of the longitudinal It is interesting to calculate the drain current of transistor
electric field, Ex , at the source end of the channel can be easily Q1 in the scattering-limited ballistic mode, with the reflection
associated with the n+ (Source)/n (Source Extension) high- coefficient of the electrons calculated as
low junction and the n (Source Extension)/p (Body) junction, 
respectively. The same identification applies to the drain end of 1 − v vT
r=  ≈ 0.18. (6)
the channel. 1 + v vT
Continuing with the asymmetric transistor [Fig. 9(b)], one
notices a third field peak at the source end of the channel, In the aforementioned equation, the average velocities v are
which is associated with the p+ (Source Halo)/p (Body) high- calculated by the drift-diffusion simulator from the local current
low junction. No additional feature is noticed in the drain densities, and vT is the thermal velocity in silicon, roughly
BULUCEA et al.: PHYSICS, TECHNOLOGY, AND MODELING OF COMPLEMENTARY ASYMMETRIC MOSFETs 2371

estimated at 1.5 × 107 cm/s, based on Assad et al.’s quantum–


mechanical calculations [49], assuming a gate length of 130 nm
for our threshold-channel transistor (matched compact mod-
eling value) and using the electron density of 5 × 1012 cm−2
calculated by the simulator.
The particular value of r quoted in (6) is calculated at a
position in the channel where the electric potential increases
by kT /q = 26 mV from the minimum value at the source-
extension/source-halo barrier, roughly corresponding to a
50-nm lateral distance from the location of that minimum
[Fig. 9(b)]. It is substantially better than r ≈ 0.4 reported before
for symmetric 0.30-μm MOSFETs [49]. This indicates a more
efficient carrier extraction at the source-channel barrier, as
expected for transistors with shorter gates and more favorable
field structure around the source-channel barrier.
Using the scattering theory equation for VD  kT /q,
one gets, with the aforementioned values and Cox = 4.7 × Fig. 10. Compared distributions of the electric potential at the surface of
10−3 F/m2 , at VG − VT = 0.2 V and VD = 1.5 V symmetric and asymmetric n-channel MOSFETs built in the empty-well archi-
tecture described in Section IV and having similar threshold voltages. Notice
  the barrier lowering induced by the doping gradient (DGIBL) in the source halo
1−r region of the asymmetric transistor.
ID1 /W = Cox vT (VG − VT ) = 98 A/m (7)
1+r
TABLE I
which represents the scattering-theory drain current of the RING OSCILLATOR STAGE DELAYS, 3-V MOSFETs
threshold transistor Q1 alone, at the specified gate overdrive
and drain voltage. This value is (1 − r)/(1 + r) = 70% of
the ballistic limit of the same transistor. However, the mea-
sured current of the composite transistor Q1 + Q2 , ID /W =
19.6 A/m, it is only 20% of the scattering-theory current of the
threshold transistor Q1 due to the effect of the main channel
transistor Q0 , which acts as a current-limiting series resistance.
In addition to limiting the current, the parasitic resistance
increases the drain saturation voltage, as already reported [49].
This is an alternate explanation of the observed increased drain
saturation voltage in long-channel asymmetric transistors (see
Section III). B. Short-Channel Case
Comparing the potential distribution in the symmetric and
asymmetric devices, a ΔΦ S = 120 mV lowering of the gate- In short-channel devices, the electron-extracting field around
controlled source-channel barrier at the surface is observed in the source-channel barrier is increased, as qualitatively dis-
the asymmetric device, which independently contributes to an cussed in Section III. Hence, the scattering theory and the
increase of the electron injection at the source. This effect is DGIBL effect can explain the observed increased drive current.
similar to the customary drain-induced barrier lowering (DIBL) The analysis is similar to the one already used for the long-
[68] but is produced by the built-in field in the halo region channel case.
rather than by the field from the drain bias. Hence, it is not
associated with a degraded output resistance of the transistor.
VII. S WITCHING S PEED P ERFORMANCE
We will call it doping-gradient-induced barrier lowering, or
DGIBL, for similarity with and to differentiate from the DIBL Having higher drive currents and transconductance (see
[68]. An explicit illustration of this effect is provided in Fig. 10, Figs. 3 and 4), the short-channel asymmetric transistors are
where the surface potential distributions in the symmetric and expected to be faster digital switchers than their symmetric
asymmetric devices are compared for transistors manufactured counterparts. This has been verified through compared ring-
in the same process and having similar threshold voltages. oscillator measurements, using 101-stage CMOS rings built
The DGIBL effect explains why asymmetric transistors, with asymmetric and symmetric 3-V transistors, each type of
while providing locally much higher dopant concentrations transistor having the minimum gate lengths allowed by the
to their body regions, can still be manufactured at the respective hot-carrier reliability requirements and widths opti-
same threshold voltage as their symmetric counterparts. mized for maximum switching speed.
Thompson et al. [74] have also observed changes in the thresh- Table I compares the stage delays and the standard de-
old voltage associated with dopant-controlled changes in the viations observed on ring oscillators built with symmetric
built-in potential barrier at the source-channel junction of short- and asymmetric 3-V transistors. The last row includes the
channel symmetric transistors using halo implants. data for even faster ring oscillators built with subminimum
2372 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 57, NO. 10, OCTOBER 2010

Fig. 12. Measured cutoff frequency fT of symmetric and asymmetric


Fig. 11. Scaling of inverter stage delays with gate lengths in CMOS ring n-channel transistors compared at the same gate overdrive voltage and similar
oscillators built with symmetric and asymmetric transistors. The two minimum gate lengths. The inset compares the maximum cutoff frequencies, fT max ,
gate lengths, L, specified on the figure are for the NMOS/PMOS transistors, of transistors with 0.4 μm gate lengths, where the value corresponding to the
respectively. The marked percentage improvements refer to the respective asymmetric transistor, not available on the test pattern, is interpolated, on the
switching speeds (∼1/τ ). trend line, between the values available at 0.3 and 0.6 μm.

(0.25 μm) gate length transistors that are still hot-carrier reli- [36], [52]. This has been primarily associated with the increased
able, in our empty-well architecture (10 years threshold lifetime transconductance of the asymmetric transistors, as already
versus 400 years for the conventionally defined minimum gate stated in Section II.
length of 0.3 μm). The ring oscillators built with the conser- An improvement of up to 20% has also been reported,
vatively defined minimum-gate-length asymmetric transistors based on device simulations, for the maximum frequency of
switch 35% faster than those built with symmetric devices. The oscillation, fmax [36].
similar comparison using subminimum gate lengths (0.25 μm) In our empty-well implementation, the measured fT im-
asymmetric devices yields a 64% speed advantage. provement in asymmetric transistors is higher than in the
In this earlier comparison, the minimum gate lengths of the aforementioned estimations due to the additional advantages
symmetric n-channel MOSFETs are slightly larger than those of a reduced channel doping, which favors increased carrier
of the asymmetric ones, i.e., 0.4 μm versus 0.3 μm. This the mobilities and reduced gate capacitances. Fig. 12 compares
due to the hot-carrier lifetime requirements, as the symmetric the silicon results of the fT measurements for asymmetric and
transistors, having no halo implants, are built on more heavily symmetric transistors at the same gate overdrive voltage. One
doped wells (see Fig. 6). Hence, a speed comparison at the can observe that, for similar gate lengths, the cutoff frequency
same gate length is in order, which is done by the extrapolation of asymmetric transistors is substantially higher than that of
of the available ring oscillator results down to 0.3 μm gate their symmetric counterparts.
lengths using Taur’s scaling equation [68]. The input and output For a comparison at exactly the same gate length, we have
capacitances of the inverter stage, Cin and Cout are assumed chosen L = 0.4 μm. The maximum cutoff frequency fT max is
to be constant in the small range of gate lengths of interest. directly available at this gate length for the symmetric transistor.
The saturation current densities per unit width in each point of The corresponding value of fT max for the asymmetric transis-
the plot are obtained from 10-μm wide transistor test structures tors was interpolated on the trend line, between the available
available on the same die with the ring oscillator structure. gate lengths, as shown in the inset of Fig. 12.
This same-gate-length comparison is included in Table I and This same-gate-length comparison results in a 1.5 times
shown in Fig. 11, revealing a 24% switching speed advantage of higher fTmax of the asymmetric transistor. However, with their
inverter stages built with asymmetric transistors over symmetric large margin in hot-carrier reliability and drain–source leakage,
counterparts having the same gate lengths. Table I and the chart the asymmetric transistors can be built at substantially smaller
in Fig. 11 also include the ring oscillator data for the high- gate lengths than 0.4 μm and even than their nominal minimum
speed asymmetric MOSFETs built at subnominal gate lengths gate length of 0.3 μm. For example, at 0.25 μm gate lengths,
that still do not compromise the hot-carrier reliability. In this the asymmetric transistors are more than two times faster than
case, the speed improvement over the ring oscillator built with 0.4 μm symmetric counterparts, while still remaining compara-
symmetric transistors is 51%. ble in hot-carrier lifetime and drain–source leakage.
Even more important than the aforementioned is the compar-
ison at low currents in the subthreshold regime. From Fig. 12,
VIII. S MALL -S IGNAL S PEED P ERFORMANCE
one can observe a 3.5 times increased cutoff frequency at
The small-signal cutoff frequency of transistors with channel zero gate overdrive. This makes the asymmetric transistors
asymmetry has been analyzed in detail by device simulations particularly attractive in low-power high-frequency applica-
that estimate up to 36% improvement in comparison with tions, where they can compete with SiGe-base bipolar junction
symmetric transistors of otherwise similar construction [33], transistors designed for the same voltage range.
BULUCEA et al.: PHYSICS, TECHNOLOGY, AND MODELING OF COMPLEMENTARY ASYMMETRIC MOSFETs 2373

IX. PARAMETER VARIABILITY, M ATCHING , AND N OISE


The halo implants, which are unavoidable in symmetric
MOSFETs beyond the 0.25 μm lithography node, are generally
known to cause higher variability and worse matching of elec-
trical properties of respective devices in comparison with those
of nonhalo counterparts [75]–[79]. They are also known to be
noisier [4]. The asymmetric transistors, having strong halos at
the source, suffer from the same limitations.
In minimum-size symmetric halo transistors, the statistical
variations of the doping impurity positioning in the halo im-
plants have been reported to account for about 15% increased
variability of the threshold voltage in comparison with devices
having ideally contained dopants [75], [76]. More recently,
variability in the 20% range was reported for heavily doped
long-channel symmetric halo transistors [78]. Fig. 13. Laterally dithered representations of threshold voltages for
minimum-size symmetric and asymmetric complementary MOSFETs. The
The increased variability of the transistors containing halo standard deviations for the asymmetric transistors are more than double with
implants, in general, is believed to be due to the contribution reference to their symmetric counterparts, for both types of transistors, when
of the DGIBL effect, which can be qualitatively explained as compared at similar threshold voltages.
follows, based on our device physics analysis in Section VI.
The high gradient of the doping concentration in the source
halo region, in the vicinity of the source-channel junction, re-
sponsible for the DGIBL effect, is the result of the combination
of very abrupt impurity profiles of opposite type in the source
extension and halo regions. As such, it has increased sensitivity
to the fluctuating details of the competing doping profiles
involved such as the implant conditions, silicon point defects,
stress effects and the statistical variations of the impurity posi-
tioning in the confined space of the source extension and halo
regions. Moreover, the angled implants used in halo transistors
involve dopant variability inherently associated with the grain-
boundary diffusion through the edges of the polysilicon gates.
These multiple sources of variability result in increased lot-to-
Fig. 14. Compared spectral densities of the drain current noise of asymmetric,
lot, die-to-die, and even same-die (mismatching) threshold volt- symmetric, and reversely biased asymmetric n-channel MOSFETs at W =
age variability in transistors having halo implants in comparison 10 μm, ID = 80 μA, measured at f = 145 Hz as functions of gate length,
to nonhalo transistors. L (median values of nine devices from one typical wafer).
The increased variability of the MOSFETs containing halo edges, excessive implant damage, etc. These requirements are
implants is also present in asymmetric MOSFETs. On one particularly critical in analog applications.
hand, the variability is expected to increase because, in order In cases when asymmetric MOSFETs are integrated for their
to set a given threshold voltage, the source-only halo implants voltage gain and speed advantages, symmetric counterparts
used in asymmetric MOSFETs have higher doses that those without halo implants need to be concurrently available for use
used in symmetric devices with the same threshold voltage. On in circuits where variability and matching are critical.
the other hand, the variability is expected to decrease, as there is The spectral densities of the 1/f (flicker) noise of our
no halo at the drain, thus eliminating one source of variability. symmetric and asymmetric MOSFETs are plotted against the
In our minimum-size asymmetric transistors, the variability gate length in Fig. 14, along with the same property of the
of the absolute threshold voltage measured on matching pairs reversely biased asymmetric devices. It is important to observe
was comparable to the figures reported before for symmetric that, when biased normally, the asymmetric devices have 10 to
halo transistors. However, it is important to observe that sym- 100 larger 1/f noise in comparison with symmetric transistors.
metric transistors not using halo implants, manufactured in the Also, unlike in their symmetric counterparts, the 1/f noise is
same process for the same threshold voltage, measured down to not scalable with the gate length.
two times smaller variability, as shown in Fig. 13. Contrary to the aforementioned, when biased reversely, the
Device variability and poor matching pose considerable chal- same asymmetric transistors are well behaved, i.e., their noise
lenges to the integration of transistors containing halo implants spectral density follows the 1/(L)2 trend of the symmetric
in general. In addition to precise control of the implantation transistors [80].2 One can see that the roughly constant value
steps, the spacer formation unit process and the processing
sequence around the application of the graded channel implant 2 This experiment is cited here only in support of the physical reasoning on
and LDD extension must be optimized to minimize variability the origin of the increased noise in asymmetric transistors. Otherwise, using
inducing effects, such as undesirable implants through spacer reversely biased asymmetric transistors is not recommended.
2374 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 57, NO. 10, OCTOBER 2010

Fig. 15. Measured and modeled output characteristics of n- and p-channel Fig. 16. Normalized transconductance, gm . L versus VGS for 10-μm-wide
transistors with with L = 5 μm and W = 10 μm, at VB = 0 V. 3-V asymmetric transistors of short and long channels (L = 0.3 and 1 μm,
respectively) at VD = 0.1 V, VB = 0 V.

of the increased spectral density measured in normally biased


asymmetric transistors corresponds to the extrapolated spectral
density of symmetric transistors at L ≈ 0.13 μm, which is the
gate length of the threshold transistor Q1 in the first-order
transistor model shown in Fig. 2. This is a strong indication
that the observed nonscalable noise of the asymmetric transistor
comes from the threshold transistor Q1 .
As a consequence of their increased 1/f noise, asymmetric
transistors are not recommended in circuits where noise is
critical. Symmetric counterparts should be used instead.

X. C OMPACT M ODELING OF A SYMMETRIC MOSFETs


Accurate modeling of the asymmetric MOSFETs is of para-
mount importance in enabling the design community to lever-
age their superior analog performance. The unavailability of Fig. 17. Maximum voltage gain, gm . rout , versus channel length, L, for
10-μm-wide 3-V asymmetric transistors at two different gate overdrive volt-
commercial compact models and the challenges associated ages, and VD = 1.5 V, VB = 0 V.
with in-house modeling have actually contributed to the rather
limited use of asymmetric transistors in the industry versus their
superior electrical capabilities. Fig. 17 shows the scaling of the maximum voltage gain,
Based on the physical understanding of the device physics gm . rout , with the channel length L, along with the expected
presented in Section VI, we developed a subcircuit model based spread, at two different gate overdrive voltages, based on the
on two BSIM3 [81] transistors, Q0 and Q1 , connected in series fast and slow corner models. The larger variability in the longer
(Fig. 2). The threshold transistor Q1 controls the subthreshold channel devices (L > 1 μm) is due to the more pronouncedly
and near-threshold regions, while the main channel transistor peaked transconductance and higher output resistance, both of
Q0 controls the saturation region of the I–V characteristic. which are inherently more difficult to control and characterize.
Scaling across geometries requires the use of binning pa- The plot also reveals that the maximum voltage gain actually
rameters, which brings further challenges, such as the need decreases for very long channel devices (L > 1 μm), which
to minimize discontinuities between adjacent bins and te- restricts the useful high range of device sizes.
dious generation of the corner models. Yet, good model ac- Fig. 18 shows the measured and modeled values of Cgs
curacy can be achieved, as shown in Fig. 15 for the output and Cgd in their dependence of the gate voltage, where the
characteristics. rising portion of the Cgd plot appears to be shifted toward
The transconductance is modeled for the minimum channel lower voltages and to peak at some greater value than the usual
length transistor (L = 0.3 μm) and for a long-channel device 1/2W L Cox of a symmetric transistor.
(L = 1 μm), as shown in Fig. 16. Notice that the transconduc- While the rather unusual behavior of Cgd has been analyzed
tance dependence on the gate voltage is more sharply peaked in in great detail in the compact modeling literature [82]–[84],
longer channel transistors, which calls for a design tradeoff be- a simple physical model of it is still desirable. The physi-
tween exploiting the increased transconductance and reducing cal structure of the asymmetric transistor (Fig. 1), combined
the biasing sensivity of the circuit performance. with the two-transistor model (Fig. 2), invite a straightforward
BULUCEA et al.: PHYSICS, TECHNOLOGY, AND MODELING OF COMPLEMENTARY ASYMMETRIC MOSFETs 2375

XI. E XTENDED -D RAIN MOSFETs IN E MPTY-W ELL


I MPLEMENTATION (“XD-MOS”)
In addition to the primary and secondary voltage transistors,
e.g., transistors biased from 1.2- and 3.0-V power supplies,
analog and mixed-signal circuits require transistors capable to
sustain a third “high” voltage, e.g., 12 V. These transistors
are typically needed in EEPROM devices, high-voltage output
buffers, display drivers, etc.
Higher voltage transistors can be integrated either as true
high-voltage laterally diffused MOSFETs (LDMOSFETs),
with the addition of a body mask and implant for each type
of transistor and with the drain-side end of the gate termi-
nating on the isolation oxide [85]–[95] or as “free” high-
voltage extended-drain MOSFETs, also called drain-extended
Fig. 18. Normalized gate capacitances Cgd and Cgs versus the gate voltage MOSFETs, using the p-well as the body region and the n-well
VGS at VDS = 0 V for a long-channel (L = 1 μm) asymmetric transistor. The as a drain extension (n-channel MOSFETs) [96]–[105].
measured Cgd rises to more than 1/2 W L Cox at gate voltages between the
threshold voltages of the component transistors Q0 (threshold) and Q1 (main). Since the terminology and description of these transis-
tors vary and are somewhat inconsistent, we will use the
high-voltage LDMOSFET for the former and high-voltage
physical explanation of the capacitance characteristics, as de- extended-drain MOSFET for the latter. Of the two varieties, the
scribed later. high-voltage extended-drain MOSFET is preferred in VLSI
circuits due to cost advantages. The high-voltage extended-
1) When the gate is biased negatively, below the threshold drain MOSFET built in the empty-well architecture will
voltages of both transistors Q0 and Q1 , the source and hereafter be referred to as high-voltage XD-MOSFET, for dif-
drain regions and their extensions provide the lower ferentiation from the one built with conventional (i.e., filled)
plates of the gate-to-source and gate-to-drain capacitors, wells.
respectively, thus reducing Cgs and Cgd to their respec- A reported difficulty with the high-voltage extended-drain
tive overlap plus fringing capacitance values. MOSFETs is their susceptibility to hot-carrier degradation
2) As the gate voltage is increased to voltages still below the [106]–[109] because the electric field strength is at its maxi-
threshold voltage of the asymmetric transistor, the chan- mum near the surface, as shown in Fig. 19(a) for an n-channel
nel transistor Q0 gradually turns on, while the threshold device.
transistor Q1 remains off. This situation can be visualized The device architecture requirement for a hot-carrier-
as having the lower plate of the gate-to-drain capacitor reliable extended-drain MOSFET is that the location of the
extended under the gate. In long-channel transistors, the
maximum impact ionization is a few carrier mean-free paths
length of the channel transistor takes up almost the entire
below the surface, so that the carriers generated by impact
length of the transistor, which makes Cgd approach the
ionization cannot make it to the silicon/silicon dioxide interface
total gate-to-channel capacitance, W L Cox .
[110]–[113]. Specifically, the drain breakdown voltage and
3) When the threshold voltage of the threshold transistor Q1
the hot-carrier lifetime of n-channel MOSFETs are critically
is reached and exceeded, the lower plate of the gate-to-
dependent on the gate overlap of drain, LOV , the optimum
source capacitor also extends into the channel, increasing
value of which has to be determined experimentally or by well
Cgs . Eventually, the channel becomes continuous from
calibrated simulations.
the source to the drain and the charge in the channel is
Even if precisely optimized, transistors built in the conven-
partitioned between the source and drain at nearly equal
tional extended-drain architecture remain sensitive to process
values of Cgd and Cgs at 1/2 W L Cox , as experimentally
variations. The location of the maximum electric field can
observed.
easily move closer to the surface, thus exposing them to hot-
The subcircuit model is also able to capture other carrier degradation.
particularities of the asymmetric devices, such as the larger The empty-well architecture used in our preferred imple-
saturation voltage in long channel devices (see Section III), mentation of the asymmetric MOSFET design is ideally suited
the independence of the 1/f noise of the channel length (see for building complementary high-voltage transistors that are
Section VIII), etc. naturally hot-carrier reliable. This stems from the electric field
Finally, additional efforts have been devoted to extracting strength being highest below the surface, where the retrograde
operating point information for the whole transistor, knowing n- and p-wells are at their peak concentrations. Due to the
that commercial simulators provide this information only for retrograde nature of the p- and n-type wells, the depletion
the individual transistors of the subcircuit model. The anno- region extends more freely below the silicon/silicon dioxide
tation on the schematic includes analytical expressions for interface, in comparison with the situation in the conventional
the transistor transconductance, output conductance, and the devices, providing field relief at the surface, as shown in
saturation voltage VDsat . Fig. 19(b). P -channel devices are built similarly, with a deep
2376 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 57, NO. 10, OCTOBER 2010

Fig. 19. (a) Cross-section of an n-channel high-voltage extended-drain MOSFET implemented in the conventional architecture with the electron flow and the
depletion region structure in forward conduction at high drain voltage. Drain breakdown occurs near the surface (simulation-based line drawing). (b) Cross-section
of an n-channel high-voltage extended-drain MOSFET implemented in the empty-well architecture (XD-MOSFET) with the electron flow and the field structure
in forward conduction at high drain voltage. Drain breakdown occurs in the bulk of the silicon structure, where the doping concentrations are at or near peak
values (simulation-based line drawing).

XII. C ONCLUSION
The physics, technology, and modeling of complementary
asymmetric MOSFETs have been reviewed, covering both the
advantages (primarily switching speed, high-frequency behav-
ior, voltage gain, and reliability) and the problems (primarily
variability and noise) of this type of transistors, with the con-
clusion that both asymmetric and symmetric devices need to be
offered in the same process for uncompromised circuit design.
The essential features of the asymmetric transistors have
been illustrated with statistically validated transistor data from
a recent manufacturing implementation, where the transistors
for the secondary power supply voltage (DGO transistors)
are offered in asymmetric and symmetric constructions. This
implementation uses empty wells, i.e., retrograde wells with no
Fig. 20. Experimental output I–V characteristics of n- and p-channel high- dopant fillers at surface and source-side-only halo implants for
voltage extended-drain MOSFET having balanced threshold voltages at |VT | ≈ threshold voltage and punch-through control.
0.5 V, implemented in the empty-well architecture (XD-MOS). The availability of silicon and simulation data for asymmetric
and symmetric transistors built in the same process and having
n-well added under the structure to isolate the drain from the similar threshold voltages has allowed an in-depth comparison
substrate. of their operation leading to the observation of a DGIBL in
The output characteristics of experimental n- and p-channel asymmetric transistors. This effect contributes to the observed
XD-MOSFETs are shown in Fig. 20 for LW W = LOV = larger drain saturation current, at similar overdrive gate volt-
0.5 μm and L = 1.0 μm, where the distance between the source ages. It also increases the variability of asymmetric transistors
edge of the gate and the drain well, L, is conventionally defined and of transistors containing halo implants, in general.
as the channel length for both types of transistors. In addition, the GISL has been analyzed and shown to
Unlike the more common high-voltage extended-drain be a limiting factor to the degree of asymmetry that can be
MOSFETs, the characteristics of these transistors do not de- accepted in applications where the source and drain electrodes
pend, over the range of practical interest, on the gate over- change bias polarity and the body and source regions are not
lap parameter LOV . Their breakdown voltage is controlled permanently connected together.
by the well-to-well separation LW W , increasing as LW W is Compact modeling of asymmetric transistors has been briefly
increased. It eventually saturates at a limit value when the discussed and illustrated with compared model and silicon char-
intensification of the electric field associated with the proximity acteristics for output I–V characteristics, transconductance, out-
of the p-well vanishes and the breakdown moves up to the put conductance, maximum voltage gain, and gate capacitances.
surface. As a side benefit, the empty-well process architecture en-
The threshold voltages are fairly well balanced through the ables construction of complementary high voltage extended-
use of the same halo implants that control the threshold volt- drain MOSFETs for a third higher power supply voltage at
ages of the secondary-voltage (DGO) asymmetric MOSFETs. superior manufacturability and hot-carrier reliability compared
Further performance optimization can be obtained through to extended-drain transistors built with conventional (i.e., filled)
parameters L, LW W , and LOV . wells.
BULUCEA et al.: PHYSICS, TECHNOLOGY, AND MODELING OF COMPLEMENTARY ASYMMETRIC MOSFETs 2377

ACKNOWLEDGMENT [14] L. T. Su, J. A. Yasaitis, and D. Antoniadis, “A high-performance scalable


sub-micron MOSFET for mixed analog/digital applications,” in IEDM
The authors would like to thank Senior Vice President Tech. Dig., 1991, pp. 367–370.
M. Yegnashankaran and Vice President R. Razouk for their [15] T. Horiuchi, T. Homma, Y. Murao, and K. Okumura, “An asymmetric
sidewall process for high performance LDD MOSFETs,” IEEE Trans.
support through the process development that made this con- Electron Devices, vol. 41, no. 2, pp. 186–190, Feb. 1994.
tributing review possible. The encouragement and manage- [16] A. Hiroki, S. Odanaka, and A. Hori, “A high performance 0.1 μm
ment support of directors S. Desai, A. Shaikh, P. Chaparala, MOSFET with asymmetric channel profile,” in IEDM Tech. Dig., 1995,
pp. 439–442.
D. Boisvert, C. Joyce, L. Smith, P. Hopper, and S. Adler [17] J. P. John, V. Ilderem, C. Park, J. Teplik, K. Klein, and S. Cheng,
are also acknowledged. The process integration management “A low-voltage graded-channel MOSFET (L-GCMOS) for sub
has been provided by W. Belcher and P. Johnson, while 1-Volt microcontroller application,” in VLSI Symp. Tech. Dig., 1996,
pp. 178–179.
D. Brisbin characterized the hot-carrier reliability. The ring [18] J. A. Babcock, C. E. Gill, J. M. Ford, D. Ngo, E. Spears, J. Ma,
oscillator speed evaluations have been done by E. Mazotti, H.-B. Liang, D. J. Spooner, and S. Cheng, “1/f noise in graded-channel
with electronic testing provided by M. Poulter and M. Liu. MOSFETs for low-power low-cost RFICs,” in Proc. Annu. Device Res.
Conf., 1997, pp. 122–123.
The 3-D visualization of the highly irregular doping data from [19] S. Odanaka and A. Hiroki, “Potential design and transport property
the process simulation has been provided by E. Russell. The of 0.1-μm MOSFET with asymmetric channel profile,” IEEE Trans.
authors would also like to thank chief technologist D. Archer Electron Devices, vol. 44, no. 4, pp. 595–600, Apr. 1997.
[20] J. Ma, H.-B. Liang, R. A. Pryorm, S. Cheng, M. H. Kaneshiro,
for introducing them the device requirements of the analog C. S. Kyono, and K. Papworth, “Graded-channel MOSFET
designers. Initial guidance received from Professor Y. Taur (GCMOSFET) for high performance, low voltage DSP applications,”
is respectfully acknowledged along with help received from IEEE Trans. Very Large Scale Integr. (VLSI) Syst., vol. 5, no. 4,
pp. 352–358, Dec. 1997.
interns K. Banoo (Purdue University) and W. Wang (University [21] B. Cheng, V. R. Rao, B. Ikegami, and J. C. S. Woo, “Realization of
of California, San Diego). sub-100 nm asymmetric channel MOSFET with excellent short-channel
performance and reliability,” in Proc. ESSDERC Tech. Dig., 1998,
pp. 520–523.
[22] H. Shin and S. Lee, “An 0.1-μm asymmetric halo by large-angle-
R EFERENCES tilt implant (AHLATI) MOSFET for high performance and reliabil-
[1] D. Buss, B. L. Evans, J. Bellay, W. Krenik, B. Haroun, D. Leipold, ity,” IEEE Trans. Electron Devices, vol. 46, no. 4, pp. 820–822,
J.-Y. Yand, and T. Moise, “SOC CMOS technology for personal Internet Apr. 1999.
products,” IEEE Trans. Electron Devices, vol. 50, no. 3, pp. 546–556, [23] B. Cheng, A. Inani, V. R. Rao, and J. C. S. Woo, “Channel engineering
Mar. 2003. for high-speed sub-1.0 V power supply deep submicron CMOS,” in VLSI
[2] K. Kuhn, R. Basco, D. Becher, M. Hattendorf, P. Packan, I. Post, Symp. Tech. Dig., 1999, pp. 69–70.
P. Vandervoorn, and I. Young, “A comparison of state-of-the-art NMOS [24] M. Stockinger, A. Wild, and S. Selberherr, “Drive performance of an
and SiGe HBT devices for analog/mixed-signal/RF circuit applications,” asymmetric MOSFET structure: The peak device,” Microelectron. J.,
in VLSI Symp. Tech. Dig., 2004, pp. 224–225. vol. 30, no. 3, pp. 229–233, Mar. 1999.
[3] B. Murmann, P. Nikaeen, D. J. Connelly, and R. W. Dutton, “Im- [25] A. Blaum, J. Victory, and C. McAndrew, “A simple physical extraction
pact of scaling on analog performance and associated modeling method for RD –RS of asymmetric MOSFETs,” in Proc. IEEE Conf.
needs,” IEEE Trans. Electron Devices, vol. 53, no. 9, pp. 2160–2167, Microelectron. Test Struct., Mar. 1999, vol. 12, pp. 141–146.
Sep. 2006. [26] S. Mahapatra, V. R. Rao, C. D. Parikh, J. Vasi, B. Cheng, and
[4] F. Boeuf, “Outlook for 32/28/22 nm manufacturing—Future SoC J. C. S. Woo, “A study of 100 nm channel length asymmetric
products and solutions,” in VLSI Symp. Tech. Dig.—Short Course, 2009, channel MOSFET by using charge pumping,” Microelectron. Eng.,
pp. 3–51. vol. 48, no. 1–4, pp. 193–196, Sep. 1999.
[5] Y. Tarui, Y. Hayashi, and T. Sekigawa, “Diffused self-aligned MOST: [27] B. Cheng, V. R. Rao, and J. C. S. Woo, “Exploration of velocity over-
A new approach for high-speed device,” in Proc. 1st Conf. Solid-State shoot in a high-performance deep sub-0.1-μm SOI MOSFET with asym-
Devices, Tokyo, Japan, 1994, pp. 617–620. metric channel profile,” IEEE Electron Device Lett., vol. 20, no. 10,
[6] S. M. Sze and K. K. Ng, Physics of Semiconductor Devices, 3rd ed. pp. 538–540, Oct. 1999.
New York: Wiley-Interscience, 2007, ch. 6. [28] T. Ghani, K. Mistry, P. Packan, M. Armstrong, S. Thompson, S. Tyagi,
[7] D. Ueda, H. Takagi, and K. Kano, “A new vertical power MOSFET and M. Bohr, “Asymmetric source/drain transistor structure for high
structure with extremely reduced on-resistance,” IEEE Trans. Electron performance sub-50 nm gate length CMOS devices,” in VLSI Symp.
Devices, vol. ED-32, no. 1, pp. 2–6, Jan. 1985. Tech. Dig., 2001, pp. 17–18.
[8] C. Bulucea and R. Rossen, “Trench DMOS transistor technology for [29] C.-S. Choi, K.-W. Kim, and W.-Y. Choi, “Design and analysis of a new
high-current (100 A range) switching,” Solid State Electron., vol. 34, self-aligned asymmetric structure for deep sub-micrometer MOSFET,”
no. 5, pp. 493–507, May 1991. Solid State Electron., vol. 45, no. 9, pp. 1673–1678, Sep. 2001.
[9] T. Dyer, J. McGinty, A. Strachan, and C. Bulucea, “Monolithic inte- [30] D. G. Borse, K. N. M. Rani, N. K. Jha, A. N. Chandorkar,
gration of trench vertical DMOS (VDMOS) power transistors into a J. Vasi, V. R. Rao, B. Cheng, and J. C. S. Woo, “Optimization and
BCD process,” in Proc. 17th Int. Symp. Power Semicond. Devices ICs, realization of sub-100-nm channel length single halo p-MOSFETs,”
Santa Barbara, CA, 2005, pp. 47–50. IEEE Trans. Electron Devices, vol. 49, no. 6, pp. 1077–1079,
[10] F. Bauer, S. C. Jain, J. Korec, V. Lauer, M. Offenberg, and P. Balk, Jun. 2002.
“Incompatibility of requirements for optimizing short channel behaviour [31] H. Deshpande, B. Cheng, and J. Woo, “Channel engineering for analog
and long-term stability in MOSFETs,” Solid State Electron., vol. 31, device design in deep submicron CMOS technology for system on chip
no. 1, pp. 27–33, Jan. 1988. application,” IEEE Trans. Electron Devices, vol. 49, no. 9, pp. 1558–
[11] T. N. Buti, S. Ogura, N. Rovedo, K. Tobimatsu, and C. F. Codella, 1564, Sep. 2002.
“Asymmetrical halo source GOLD drain (HS-GOLD) deep sub-half- [32] T. B. Hook, J. S. Brown, M. Breitwisch, D. Hoyniak, and
micron n-MOSFET design for reliability and performance,” in IEDM R. Mann, “High-performance logic and high-gain analog CMOS tran-
Tech. Dig., 1989, pp. 617–620. sistors formed by a shadow-mask technique with a single implant
[12] T. N. Buti, S. Ogura, N. Rovedo, and K. Tobimatsu, “A new asymmet- step,” IEEE Trans. Electron Devices, vol. 49, no. 9, pp. 1623–1627,
rical halo source GOLD drain (HS-GOLD) deep sub-half-micrometer Sep. 2002.
n-MOSFET design for reliability and performance,” IEEE Trans. [33] T. Ewert, L. Vestling, and J. Olsson, “Investigation of the electrical
Electron Devices, vol. 38, no. 8, pp. 1757–1764, Aug. 1991. behavior of an asymmetric MOSFET,” Microelectron. Eng., vol. 65,
[13] T. Matsuki, F. Asakura, S. Saitoh, H. Matsumoto, M. Fukuma, no. 4, pp. 428–438, Apr. 2003.
and N. Kawamura, “Laterally-doped channel (LDC) structure for [34] F.-C. Wang and C. Bulucea, “Use of mask shadowing and angled implan-
sub-quarter micron MOSFETs,” in VLSI Symp. Tech. Dig., 1991, tation in fabricating asymmetrical field-effect transistors for alleviating
pp. 113–114. short-channel effects,” U.S. Patent 6 566 204, May 20, 2003.
2378 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 57, NO. 10, OCTOBER 2010

[35] H.-S. Huang, S.-Y. Chen, Z.-W. Jhou, and Y.-D. Zeng, “Comparison [53] J. F. Chen, J. Tao, P. Fang, and C. Hu, “Performance and reliabil-
of single and double pocket devices on deep submicron n-MOSFET ity comparison between asymmetric and symmetric LDD devices and
for mixed signal applications,” in Proc. Int. Symp. Nano Sci. Technol., logic gates,” IEEE J. Solid-State Circuits, vol. 34, no. 3, pp. 367–371,
Taiwan, 2004, pp. 20–21. Mar. 1999.
[36] K. Narasimhulu, M. P. Desai, S. G. Narendra, and V. R. Rao, “The [54] F. Chai, C. Kyono, V. Ilderem, M. Kaneshiro, D. Zupac, S. Bigelow,
effect of lateral asymmetric channel doping on deep submicrom- C. Ramiah, P. Dahl, R. Braithwaite, D. Morgan, S. Hildreth, and
eter transistor capacitances and its influence on RF performance,” G. Grynkewich, “Cost-effective 0.25 μm Leff BiCMOS technology
IEEE Trans. Electron Devices, vol. 51, no. 9, pp. 1416–1423, featuring graded-channel CMOS (GCMOS) and a quasi-self-aligned
Sep. 2004. (QSA) NPN for RF wireless applications,” in Proc. BCTM, 2000,
[37] A. S. Roy, Y. S. Chauhan, C. C. Enz, and J.-M. Sallese, “Noise pp. 110–113.
modeling in lateral asymmetric MOSFET,” in IEDM Tech. Dig., 2006, [55] A. Akturk, N. Goldsman, and G. Metze, “Increased CMOS inverter
pp. 191–194. switching speed with asymmetrical doping,” Solid State Electron.,
[38] J. Yuan, S. S. Tan, Y. M. Lee1, J. Kim, R. Lindsay, V. Sardesai, vol. 47, no. 2, pp. 185–192, Feb. 2003.
T. Hook, R. Amos, Z. Luo, W. Lee, S. Fang, T. Dyer, N. Rovedo, [56] K. Narasimhulu, S. G. Narendra, and V. R. Rao, “Impact of lateral
R. Stierstorfer, Z. Yang, J. Li, K. Barton, H. Ng, J. Sudijono, J. Ku, asymmetric channel doping on deep submicron mixed-signal device and
M. Hierlemann, and T. Schim, “A 45 nm low cost low power platform by circuit performance,” IEEE Trans. Electron Devices, vol. 50, no. 12,
using integrated dual-stress-liner technology,” in VLSI Symp. Tech. Dig., pp. 2481–2489, Dec. 2003.
2006, pp. 100–101. [57] J. J. Buller, R. van Bentum, J. Cheek, E. Ehrichs, M. Horstmann, and
[39] J. P. Kim, W. Y. Choi, J. Y. Song, S. W. Kim, J. D. Lee, and S. Searles, “Opportunities and challenges in asymmetric device
B.-G. Park, “Design and fabrication of asymmetric MOSFETs using implementation,” in Proc. IEEE Custom Integr. Circuits Conf., 2004,
a novel self-aligned structure,” IEEE Trans. Electron Devices, vol. 54, pp. 229–232.
no. 11, pp. 2969–2964, Nov. 2007. [58] B. Wang, H. Nguyen, J. Mavoori, A. Horch, Y. Ma, T. Humes, and
[40] S. K. H. Fung, H. C. Lo, C. F. Cheng, W. Y. Lu, K. C. Wu, R. Paulsen, “Effect of layout orientation on the performance and re-
K. H. Chen, D. H. Lee, Y. H. Liu, I. L. Wu, C. T. Li, C. H. Wu, liability of high voltage N-LDMOS in standard submicron logic STI
F. L. Hsiao, T. L. Chen, W. Y. Lien, C. H. Huang, P. W. Wang, CMOS process,” in Proc. Annu. Int. Rel. Phys. Symp., San Jose, CA,
Y. H. Chiu, L. T. Lin, K. Y. Chen, H. J. Tao, H. C. Tuan, Y. J. Mii, and 2005, pp. 654–655.
Y. C. Sun, “45 nm CMOS technology with 3X hole mobility enhance- [59] A. Bansal and K. Roy, “Asymmetric halo CMOSFET to reduce static
ment and asymmetric transistor for high performance CPU application,” power dissipation with improved performance,” IEEE Trans. Electron
in VLSI Symp. Tech. Dig., 2007, pp. 1035–1037. Devices, vol. 52, no. 3, pp. 397–405, Mar. 2005.
[41] G. H. See, X. Zhou, K. Chandarasekaram, S. B. Chiah, Z. Zhu, C. Wei, [60] D. V. Kumar, K. Narasimhulu, P. S. Reddy, M. Shojaei-Baghini,
S. Lin, G. Zhu, and G. H. Lim, “A compact model satisfying D. K. Shrama, M. B. Patil, and V. R. Rao, “Evaluation of the impact of
Gummel symmetry in higher order derivatives and applicable to asym- layout on device and analog circuit performance with lateral asymmet-
metric MOSFETs,” IEEE Trans. Electron Devices, vol. 55, no. 2, ric channel MOSFETs,” IEEE Trans. Electron Devices, vol. 52, no. 7,
pp. 624–631, Feb. 2008. pp. 1603–1609, Jul. 2005.
[42] M. A. Pavanello, J. A. Martino, V. Dessard, and D. Flandre, “Ana- [61] K. Narasimhulu, I. V. Setty, and V. R. Rao, “The effect of single-halo
log performance and application of graded-channel fully depleted SOI doping on the low-frequency noise performance of deep submicrometer
MOSFETs,” Solid State Electron., vol. 44, no. 7, pp. 1219–1222, MOSFETs,” IEEE Electron Device Lett., vol. 27, no. 12, pp. 995–997,
Jul. 2000. Dec. 2006.
[43] M. A. Pavanello, J. A. Martino, and D. Flandre, “Analog circuit de- [62] J.-J. Kim, A. Bansal, R. Rao, S.-H. Lo, and C.-T. Chuang, “Relaxing
sign using graded-channel silicon-on-insulator nMOSFETs,” Solid State conflict between read stability and writability in 6T SRAM cell using
Electron., vol. 46, no. 8, pp. 1215–1225, Aug. 2002. asymmetric transistors,” IEEE Electron Device Lett., vol. 30, no. 8,
[44] V. Kilchytska, A. Neve, L. Vancaillie, D. Levacq, S. Adriansen, pp. 852–854, Aug. 2009.
H. Van Meer, K. De Meyer, C. Raynaud, M. Dehan, J.-P. Raskin, and [63] K. Nii, M. Yabuchi, Y. Tsukamoto, Y. Hirano, and T. Iwamatsu, “A 0.5 V
D. Flandre, “Influence of device engineering on the analog and RF 100 MHz PD-SOI SRAM with enhanced read stability and write margin
performance of SOI MOSFETs,” IEEE Trans. Electron Devices, vol. 50, by asymmetric MOSFET and forward body bias,” in Proc. Int. Solid-
no. 3, pp. 577–588, Mar. 2003. State Circuits Conf., 2010, pp. 356–357.
[45] A. Cerdeira, M. A. Alemán, M. A. Pavanello, J. A. Martino, [64] K. K. O and J. Yasaitis, “Integration of two gate oxide thicknesses
L. Vancaillie, and D. Flandre, “Advantages of the graded-channel SOI in a 0.6-μm dual voltage mixed signal CMOS process,” IEEE Trans.
FD MOSFET for application as a quasi-linear resistor,” IEEE Trans. Electron Devices, vol. 42, no. 1, pp. 190–192, Jan. 1995.
Electron Devices, vol. 52, no. 5, pp. 967–972, May 2005. [65] D. Buss, “Device issues in the integration of analog/RF functions in deep
[46] J. Hoentschel, A. Wei, M. Wiatr, A. Gehring, T. Scheiper, R. Mulfinger, submicron digital CMOS,” in IEDM Tech. Dig., 1999, pp. 423–426.
T. Feudel, T. Lingner, A. Poock, S. Muehle, C. Krueger, T. Herrmann, [66] C. Bulucea, F.-C. Wang, and P. Chaparala, “Field-effect transistors
W. Klix, R. Stenzel, R. Stephan, P. Huebler, T. Kammler, P. Shi, M. Raab, for alleviating short-channel effects,” U.S. Patent 6 548 842, Apr. 15,
D. Greenlaw, and M. Horstmann, “Implementation and optimization 2003.
of asymmetric transistors in advanced SOI CMOS technologies [67] A. S. Grove, Physics and Technology of Semiconductor Devices.
for high performance microprocessors,” in IEDM Tech. Dig., 2008, New York: Wiley, 1968.
pp. 649–652. [68] Y. Taur and T. K. Ning, Fundamentals of Modern VLSI Devices, 2nd ed.
[47] M. Lundstrom, “Scattering theory of the short-channel MOSFET,” in Cambridge, U.K.: Cambridge Univ. Press, 2009.
IEDM Tech. Dig., 1996, pp. 387–390. [69] M. Langenbuch, R. Bottini, M. E. Vitali, and G. Ghidini, “In situ
[48] M. Lundstrom, “Elementary scattering theory of the Si MOSFET,” IEEE steam generation (ISSG) versus standard steam technology: impact on
Electron Device Lett., vol. 18, no. 7, pp. 361–363, Jul. 1997. oxide reliability,” Microelectron. Reliab., vol. 45, no. 5/6, pp. 875–878,
[49] F. Assad, Z. Ren, S. Datta, M. Lundstrom, and P. Bendix, “Performance May/Jun. 2005.
limits of silicon MOSFETs,” in IEDM Tech. Dig., 1999, pp. 547–550. [70] K. K. Young, S. Y. Wu, C. C. Wu, C. H. Wang, C. T. Lin, J. Y. Cheng,
[50] G. Yeap and A. Wild, “Introduction to low-power VLSI design,” Int. J. M. Chiang, S. H. Chen, T. C. Lo, Y. S. Chen, J. H. Chen, L. J. Chen,
High-Speed Electron. Syst., vol. 7, no. 2, pp. 223–248, 1996. S. Y. Hou, J. J. Law, T. E. Chang, C. S. Hou, J. Shih, S. M. Jeng,
[51] A. Wild, J. Quigley, J. Feddeler, S. Ledford, J. Caravella, F. Shapiro, H. C. Hsieh, Y. Ku, T. Yen, H. Tao, L. C. Chao, S. Shue, S. M. Jang,
B. Gilsdorf, M. Hong, D. Mietus, R. B. Davies, R. Parmar, J.-D. Tai, T. C. Ong, C. H. Yu, M. S. Liang, C. H. Diaz, and J. Y. C. Sun, “A
J. Thomas, J. Quarberg, E. Hartung, T. Sawan, J. H. Carlquist, 0.13 μm CMOS technology with 193 nm lithography and Cu/low-k for
K. Papworth, J. Buxo, M. Schriber, M. Berens, D. Warren, B. Smith, high performance applications,” in IEDM Tech. Dig., 2000, pp. 563–566.
M. Mazuelos, K. Bass, and L. Layton, “A 0.9 V microcontroller for [71] J. Sall, L. Creighton, and A. Lehman, JMP Start Statistics: A Guide
portable applications,” IEEE J. Solid-State Circuits, vol. 32, no. 7, to Statistics and Data Analysis Using JMP, 4th ed. Cary, NC: SAS
pp. 1049–1055, Jul. 1997. Publishing, 2007.
[52] D. Lamey, T. Mackie, H.-B. Liang, J. Ma, G. Robert, C. Jasper, D. Ngo, [72] Synopsis, Inc., Taurus TSuprem-4 User Guide, 2004.
K. Papworth, S. Cheng, C. Wilcock, R. Gurrola, E. Spears, and B. Yeung, [73] Synopsis, Inc., Taurus Medici User Guide, 2004.
“Improving manufacturability of an RF graded channel CMOS process [74] S. Thompson, P. Packan, and M. Bohr (1998). MOS scaling: Transistor
for wireless applications,” in Proc. SPIE Microelectron. Device Technol. challenges for the 21st century. Intel Technol. J. [Online]. 2(3), pp. 1–2.
Conf., 1998, pp. 147–155. Available: http://www.intel.com/technology/itj/archive/1998.htm
BULUCEA et al.: PHYSICS, TECHNOLOGY, AND MODELING OF COMPLEMENTARY ASYMMETRIC MOSFETs 2379

[75] T. Tanaka, T. Usuki, T. Futatsugi, Y. Momiyama, and T. Sugii, “Vth [100] T. Elfand, T. Keller, S. Keller, and J. Rodriguez, “Optimized
fluctuation induced by statistical variation of pocket dopant profile,” in complementary 40 V power LDMOS-FETs use existing fabricating
IEDM Tech. Dig., 2000, pp. 271–274. steps in submicron CMOS technology,” in IEDM Tech. Dig., 1994,
[76] T. Tanaka, T. Usuki, Y. Momiyama, and T. Sugii, “Direct measurement pp. 399–402.
of Vth fluctuation caused by impurity positioning,” in VLSI Symp. Tech. [101] Y. Q. Li, T. Salama, M. Seufert, P. Schvan, and M. King, “Inte-
Dig., 2000, pp. 136–137. gration of high-voltage NMOS devices into a submicron BiCMOS
[77] K. Kuhn et al., “Managing process variations in Intel’s 45 nm CMOS process using simple structural changes,” in IEDM Tech. Dig., 1994,
technology,” Intel Tech. J., vol. 12, no. 2, pp. 93–109, Jun. 2008. pp. 403–406.
[78] A. Cathignol, S. Bordez, A. Cros, K. Rochereau, and G. Ghibaudo, “Ab- [102] P. M. Santos, A. P. Casimiro, M. I. C. Simas, and M. Lança, “CMOS
normally high local electrical fluctuations in heavily pocket-implanted compatible HV gate-shifted LDD-NMOS,” IEEE Trans. Electron
bulk long MOSFET,” Solid State Electron., vol. 53, no. 2, pp. 127–133, Devices, vol. 48, no. 5, pp. 1013–1015, May 2001.
Feb. 2009. [103] J. C. Mitros, C. Tsai, H. Shichijo, K. Kunz, A. Morton, D. Goodpaster,
[79] J. A. Croon, E. Augendre, S. Decoutere, W. Sansen, and H. E. Maes, D. Mosher, and T. R. Efland, “High-voltage drain extended MOS transis-
“Influence of doping profile and halo implantation on the threshold tors for 0.18-μm logic CMOS process,” IEEE Trans. Electron Devices,
voltage mismatch of a 0.13 μm CMOS technology,” in Proc. ESSDERC, vol. 48, no. 8, pp. 1751–1755, Aug. 2001.
2002, pp. 579–582. [104] P. Moens, B. Vlachakis, F. Bauwens, and L. De Schepper, “Spatial distri-
[80] A. van der Ziel, Noise in Solid-State Devices and Circuits. New York: bution of interface traps in DeMOS transistors,” IEEE Electron Device
Wiley, 1986. Lett., vol. 25, no. 8, pp. 577–579, Aug. 2004.
[81] Y. Cheng and C. Hu, MOSFET Modeling & BSIM3 User’s Guide. [105] P. Moens, F. Bauwens, and M. Tack, “Electron trapping and interface
Norwell, MA: Kluwer, 1999. trap generation in drain extended PMOS transistors,” in Proc. IEEE Int.
[82] J. Jang, T. Arnborg, Z. Yu, and R. W. Dutton, “Circuit model for Rel. Phys. Symp., 2005, pp. 555–559.
power LDMOS including quasi-saturation,” in Proc. SISPAD, 1999, [106] J. F. Chen, K.-M. Wu, K.-W. Lin, Y.-K. Su, and S. L. Hsu, “Hot-carrier
pp. 15–18. reliability in submicrometer 40 V LDMOS transistors with thick gate
[83] A. C. T. Aarts, R. van der Hout, J. C. Y. Paasschens, A. J. Scholten, oxide,” in Proc. Int. Rel. Phys. Symp., 2005, pp. 560–564.
M. Willensen, and D. B. M. Klaassen, “New fundamental insights into [107] B. Wang, H. Nguyen, J. Mavoori, A. Horch, Y. Ma, T. Humes, and
capacitance modeling of laterally nonuniform MOS devices,” IEEE R. Paulsen, “Effect of layout orientation on the performance and reliabil-
Trans. Electron Devices, vol. 53, no. 2, pp. 270–277, Feb. 2006. ity of high voltage N-LDMOS in standard submicron logic STI CMOS
[84] A. S. Roy, Y. S. Chauhan, J.-M. Sallese, C. C. Enz, A. M. Ionescu, and process,” in Proc. Annu. Int. Rel. Symp., 2005, pp. 654–655.
M. Declercq, “Partitioning scheme in lateral asymmetric MOST,” in [108] D. Brisbin, A. Strachan, and P. Chaparala, “Optimizing the hot carrier re-
Proc. ESSDERC, 2006, pp. 307–310. liability of N-LDMOS transistor arrays,” Microelectron. Reliab., vol. 45,
[85] S. C. Sun and J. D. Plummer, “A monolithic 200-V CMOS analog no. 7/8, pp. 1021–1032, Jul./Aug. 2005.
switch,” IEEE J. Solid-State Circuits, vol. SC-11, no. 6, pp. 809–817, [109] R. A. Bianchi, F. Monsieur, F. Blanchet, C. Raynaud, and O. Noblanc,
Dec. 1977. “High voltage devices integration into advanced CMOS technologies,”
[86] S. C. Sun and J. D. Plummer, “Modeling of the on-resistance of LDMOS, in IEDM Tech. Dig., 2008, pp. 137–140.
VDMOS, and VMOS power transistors,” IEEE Trans. Electron Devices, [110] C. Bulucea, “Avalanche injection into the oxide in silicon gate-controlled
vol. ED-27, no. 2, pp. 356–367, Feb. 1980. devices—I. Theory,” Solid State Electron., vol. 18, no. 4, pp. 363–374,
[87] M. Amato and V. Rumennik, “Comparison of lateral and vertical DMOS Apr. 1975.
specific on-resistance,” in IEDM Tech. Dig., 1985, pp. 736–739. [111] C. Bulucea, “Avalanche injection into the oxide in silicon gate-controlled
[88] O. Kwon, T. Elfand, T. Ng, S. Malhi, R. Todd, and J. Lee, “Optimized devices—II. Experimental results,” Solid State Electron., vol. 18, no. 5,
60-V lateral DMOS devices for VLSI power applications,” in VLSI pp. 381–391, May 1975.
Symp. Tech. Dig., 1991, pp. 115–116. [112] T. H. Ning, C. M. Osburn, and H. N. Yu, “Emission probability of hot
[89] P. G. Y. Tsui, P. V. Gilbert, and S. W. Sun, “Integration of power LDMOS electrons from silicon into silicon dioxide,” J. Appl. Phys., vol. 48, no. 1,
into a low-voltage 0.5 μm BiCMOS technology,” in IEDM Tech. Dig., pp. 286–293, Jan. 1977.
1992, pp. 27–30. [113] P. E. Cottrell, R. R. Troutman, and T. H. Ning, “Hot-electron emission
[90] M. Smayling, J. Reynolds, D. Redwine, S. Keller, and G. Falessi, “A in n-channel IGFETs,” IEEE Trans. Electron Devices, vol. ED-26, no. 4,
modular merged technology process including submicron CMOS logic, pp. 520–533, Apr. 1979.
nonvolatile memories, linear functions, and power components,” in Proc.
IEEE Custom Integr. Circuits Conf., 1993, pp. 24.5.1–24.5.4.
[91] C. Duvvury, J. Rodriguez, C. Jones, and M. Smyling, “Integration for
ESD robustness of high voltage power MOSFETs,” in IEDM Tech. Dig., Constantin Bulucea (S’69–M’70–SM’88–F’04) re-
1994, pp. 407–410. ceived the M.S. and Ph.D. degrees in electrical engi-
[92] P. G. Y. Tsui, P. V. Gilbert, and S. W. Sun, “A versatile half-micron com- neering from the Polytechnic Institute of Bucharest,
plementary BiCMOS technology for microprocessor-based smart power Bucharest, Romania, in 1962 and 1974, respec-
applications,” IEEE Trans. Electron Devices, vol. 42, no. 3, pp. 564–570, tively. In 1969, he was awarded a one-year gov-
Mar. 1995. ernment scholarship for studies at the University of
[93] C. Duvvury, D. Briggs, J. Rodriguez, F. Carvajal, A. Young, and California, Berkeley, where he received the second
M. Smayling, “Efficient NPN operation in high voltage NMOSFET for M.S. degree in electrical engineering.
ESD robustness,” in IEDM Tech. Dig., 1995, pp. 345–348. From 1965 to 1986, he worked with the rapidly
[94] R. Versari, A. Pieracci, S. Manxini, C. Contiero, and B. Ricco, “Hot- growing semiconductor industry of his country, hold-
carrier reliability in LDMOS transistors,” in IEDM Tech. Dig., 1997, ing various positions from Staff Engineer to Director
pp. 371–374. of Research and Development. During that period, he published reference
[95] R. Versari and A. Pieracci, “Experimental study of hot-carrier effects contributions to the physical electronics of surface breakdown and hot-carrier-
in LDMOS transistors,” IEEE Trans. Electron Devices, vol. 46, no. 6, injection in silicon devices. In 1987, he joined Siliconix, Santa Clara, CA,
pp. 1228–1233, Jun. 1999. where he brought to completion the development of the world’s highest current
[96] W. G. Meyer, G. W. Dick, K. H. Olson, K. H. Lee, and J. A. Shimer, transistor switch of the time in trench DMOS technology. Since 1990, he has
“Integrable high voltage CMOS: Devices, process application,” in IEDM been with National Semiconductor Corporation, Santa Clara, CA, where he
Tech. Dig., 1985, pp. 732–735. currently holds a Chief Technologist position, leading IC technology develop-
[97] G. M. Dolni, O. H. Schade, Jr., B. Goldsmith, and ment programs that integrate submicrometer CMOS and bipolar devices for
L. A. Goodman, “Enhanced CMOS for analog-digital power IC analog and mixed-signal circuits/systems. He has published over 50 papers in
applications,” IEEE Trans. Electron Devices, vol. ED-33, no. 12, the areas of device physics and engineering (1962–1980), technology of power
pp. 1985–1991, Dec. 1986. DMOS transistors (1987–1990), and VLSI devices (1991 to present). In the
[98] J. Haas, K. Au, L. C. Martin, T. L. Portlock, and T. Sakurai, “High aforementioned areas, he has 30 U.S. patents, with several others pending.
voltage CMOS LCD driver using low voltage CMOS process,” in Proc. Dr. Bulucea has served one full Technical Committee term of the VLSI
Custom Integr. Circuits Conf., 1989, pp. 14.6.1–14.6.4. Technology Symposium and the Bipolar Circuits and Technology Meeting
[99] O.-K. Kwon, T. Elfand, W. T. Ng, S. Mahli, R. Todd, and J. K. Lee, and is currently the analog and mixed-signal technology Editor of the IEEE
“Optimized 60-V lateral DMOS devices for VLSI applications,” in VLSI E LECTRON D EVICES L ETTERS. He is an honorary member of the Romanian
Symp. Tech. Dig., 1991, pp. 115–116. Academy.
2380 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 57, NO. 10, OCTOBER 2010

Sandeep R. Bahl (S’84–M’93–SM’99) received the Tikno Harjono received the B.S. and M.S. degrees
B.S. degree from Rensselaer Polytechnic Institute, in electrical engineering from Santa Clara Univer-
Troy, NY, in 1985, and the M.S. and Ph.D. degrees sity, Santa Clara, CA, in 1988 and 1990, respectively.
in electrical engineering from the Massachusetts He was a Research Intern at Xerox Palo Alto
Institute of Technology, Cambridge, in 1988 and Research Center, Palo Alto, CA, while he was work-
1993, respectively. ing toward the M.S. degree. In 1990, he joined the
From 1993 to 2005, he worked with Hewlett Device Modeling Group at Signetics/Philips Semi-
Packard and Agilent Technologies in Palo Alto, CA. conductors, Sunnyvale, CA. From 1996 to 1999, he
Since 2005, he has been with National Semiconduc- worked with Sipex Corporation, Milpitas, CA. In
tor Corporation, Santa Clara, CA. He made contri- 1999, he joined the Advanced Process Technology
butions to electronic and optical devices in III–V Development Group of National Semiconductor
materials and silicon. Among these, he conducted the research phase of the Corporation, Santa Clara, CA, as a Device Engineer.
InGaP HBT, which is now used in Agilent products, and codeveloped the next- Mr. Harjono is a member of Tau Beta Pi.
generation analog CMOS process at National Semiconductor Corporation. He
has authored or coauthored 32 journal articles and conference presentations
and holds 16 patents. His current interest is the development of energy-efficient
transistor technology for power applications.
Dr. Bahl was a recipient of the STAR awards at both Hewlett-Packard and
Agilent Technologies and received the Best Student Paper Award at the Indium
Phosphide and Related Materials Conference. He is a member of Tau Beta Pi
and Eta Kappa Nu and is an officer of the Santa Clara, CA chapter of the IEEE Vijay Krishnamurthy received the B.Tech. degree
Electron Devices Society. in electrical engineering from the Indian Institute
of Technology, Chennai, India, in 1997, and the
Ph.D. degree in electrical and computer engineer-
ing from Purdue University, West Lafayette, IN, in
William D. French was born in Northern Ireland,
in 1967. He received the B.Eng. and Ph.D. degrees 2002, where he worked on design and optimization
from the Department of Electrical and Electronic of large-area high-speed MSM photodetectors on
Engineering, Queen’s University of Belfast, Belfast, intermediate temperature grown GaAs.
Northern Ireland, in 1989 and 1993, respectively. His In 2000, he was with the IBM T.J. Watson Re-
Ph.D. thesis was on the 2-D simulation of silicon-on- search Center, Yorktown Heights, NY and worked on
insulator MOSFETs. noise characterization of photoreceivers for 10-GB
He was a Research Fellow in Queen’s University Ethernet applications. Since 2004, he has been with National Semiconductor
on high-speed silicon-on-insulator bipolar transistors Corporation, Santa Clara, CA. His current work and research interests include
before joining Silvaco International, Santa Clara, CA compact modeling of DMOS/HV-MOS devices for power switching appli-
in 1995 as an Applications Engineer for Technology cations and low-frequency noise characterization and modeling of advanced
CAD (TCAD). In 2004, he joined National Semiconductor Corporation, Santa CMOS devices.
Clara, as a TCAD Device Engineer and is currently the Manager of the TCAD Dr. Krishnamurthy was the recipient of the IEEE Lasers and Electro-Optics
group. He is responsible for the process and device simulations for all process Society Graduate Student Fellowship in November 2000.
development within National Semiconductor Corporation. His interests include
submicrometer CMOS process and device physics, LDMOS design, bipolar
device physics, and RF spiral inductor simulation and design.

Jeng-Jiun Yang (M’96) received the B.S. degree in chemistry from National
Taiwan University, Taipei, Taiwan, in 1978, and the Ph.D. degree in physical Jon Tao received the B.S. and M.S. degrees in elec-
chemistry from the University of California, Los Angeles, in 1984. trical engineering from Peking University, Beijing,
In 1982, she was a Guest Scientist at the Max Plank Institut für Quantenoptik, China, in 1984 and 1987, respectively, and the Ph.D.
Garching, Germany. Between 1984 and 1987, she held postdoctoral research degree in electrical engineering from the University
positions with the Department of Chemistry, Oregon State University, Corvallis, of California, Berkeley, in 1995.
and AT&T Bell Laboratories, Murray Hill, NJ. Since 1987, she has been From 1987 to 1990, he was an Assistant Professor
working with the IC industry of the Silicon Valley, CA, as a Professional with the Institute of Microelectronics, Peking Uni-
Contributor at Integrated Device Technology, Santa Jose, CA (1987–1997), versity. From 1995 to 2006, he worked for Siliconix,
Maxim Integrated Products, Sunnyvale, CA (1997), and National Semiconduc- Advanced Micro Devices, and RF Micro Devices in
tor Corporation, Santa Clara, CA (1997 to present). Her experience in the field the fields of process technology, process integration,
includes CMOS process development, device physics, device bench/electronic device reliability, RFCMOS, and passives modeling
characterization, and ESD/latch-up engineering. and characterization. Since 2006, he has been with National Semiconductor
Corporation, Santa Clara, CA. His area of interest is active and passive
devices physics, modeling, characterization, and spiral inductors in CMOS
Pascale Francis received the M.S.E.E. and Ph.D. technologies. He has published over 40 journal and conference papers, and is
degrees in electrical engineering from the Catholic the holder of several patents related to device physics and process technologies.
University of Louvain, Louvain-la-Neuve, Belgium,
in 1991 and 1996, respectively. Her doctoral research
was on gate-all-around SOI/CMOS device character-
ization and modeling in harsh environments, such as
high temperature and irradiation.
In 1996, she was the recipient of a fellowship from
the Japanese Association of University Women and
Courtney Parker (M’91) received the B.S. and M.S.
spent three months with the NEC Research Laborato-
degrees in electrical engineering from the University
ries, Tsukuba, Japan. She then joined National Semi-
of Maine, Orono, in 1989 and 1991, respectively, and
conductor Corporation, Santa Clara, CA, where she held various positions since
1996 in process and device characterization and modeling. She is currently a the M.B.A. degree from the University of Southern
Senior Engineering Manager in the Spice Modeling group. Her interests span Maine, Portland, in 2010.
multiple areas, such as deep submicrometer CMOS for analog applications, He is currently an Integration Engineer with the
high-frequency passive devices (inductors, transformers, varactors, etc.), high- Process Development Group of National Semicon-
voltage and high-power silicon LDMOS transistors, GaN HEMTs, etc. She ductor Corporation, Santa Clara, CA, with a focus on
authored or coauthored more than 20 publications in IEEE or international submicrometer CMOS and BiCMOS technologies.
journals and conferences.

Vous aimerez peut-être aussi