IEEE

advertisement
IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 50, NO. 3, MARCH 2003
645
Transistor Design and Application Considerations for
>200-GHz SiGe HBTs
Greg Freeman, Senior Member, IEEE, Basanth Jagannathan, Member, IEEE, Shwu-Jen Jeng,
Jae-Sung Rieh, Member, IEEE, Andreas D. Stricker, David C. Ahlgren, and Seshadri Subbanna, Member, IEEE
Invited Paper
Abstract—SiGe HBT transistors achieving over 200 GHz
and
MAX are demonstrated in this paper. Techniques and trends in
SiGe HBT design are discussed. Processing techniques available to
silicon technologies are utilized to minimize parasitic resistances
and capacitances and thereby establish raw speeds exceeding III–V
devices despite the higher mobility in those materials. Higher current densities and greater avalanche currents, which are required
for establishing such high performance, are discussed as they relate to device self-heating and reliability and the degradation of
the devices. Simple circuit results are shown, demonstrating 4.2-ps
ring-oscillator delays.
Index Terms—BiCMOS integrated circuits, bipolar transistors,
heterojunctions, semiconductor devices.
I. INTRODUCTION
B
IPOLAR transistors have historically attained faster
transit times compared to silicon FET transistors. This
is in part because, compared to FET transistors, the bipolar
transistor relies on vertical layer dimensions rather than leading
edge lithographic dimensions to attain its speed performance.
Vertical transit time from the emitter, through the base, and to
the neutral collector is determined largely by the thickness of
the base layer and base-collector space-charge regions and not
by lithography. A further favorable factor in transistor design
is that the structural flexibility available to the designer in
constructing the transistor (especially in silicon technologies)
provides degrees of freedom in minimizing the parasitic access
resistances and parasitic capacitances.
This speed advantage, together with transconductance
noise, and
(i.e., current drive), high self-gain, low
matching/reproducibility, continue to make the bipolar transistor the device of choice for many demanding applications.
Forty-gigabit/s applications are now being produced utilizing
of
SiGe bipolar and BiCMOS production technologies with
120 GHz. At the same time, near-production 0.13- m CMOS
of approximately 90 GHz are used to
technologies with
figure
produce 10-Gb/s applications. The similarity of the
of merit and, at the same time, the factor of four difference in
Manuscript received May 28, 2002; revised October 3, 2002. The review of
this paper was arranged by Editor J. D. Cressler.
The authors are with the IBM Microelectronics Semiconductor Research
and Development Center, Hopewell Junction, NY 12533 USA (e-mail:
freemang@us.ibm.com).
Digital Object Identifier 10.1109/TED.2003.810467
Fig. 1. f versus I for different generations of SiGe HBT. Near
twice-minimum emitter areas are shown for each technology at 1 m , 0.26
m , and 0.18 m from bottom to top.
application-speed illustrates the strength of the bipolar transistor. Other applications, such as high-speed test equipment
and high-frequency radar systems also take advantage of the
highest speed transistors.
The high-speed performance may also be traded for lower
versus current comparison for
power. Shown in Fig. 1 is an
three different transistor generations: a 0.5- m 47-GHz SiGe
HBT, a 0.18- m 120-GHz SiGe HBT, and a 0.12- m 210-GHz
SiGe HBT. With the lateral and vertical scaling to be discussed
in this paper, each successive transistor has higher performance
at the same current compared to the one below in the chart.
Applications not requiring the higher performance of each successive transistor may instead operate the transistor at a lower
current. As one can see from the chart, a 2 higher frequency
may be traded for approximately a 10 reduction in power dissipation at the same performance. Further, an alternate transistor
may be designed with reduced collector doping only, to achieve
a similar low current , but at a more favorable collector–base
capacitance and breakdown.
With the recent demonstration of SiGe transistors operating
and
[1], it is interesting to
at greater than 200 GHz
take note that SiGe HBT speeds are now surpassing the best
0018-9383/03$17.00 © 2003 IEEE
646
IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 50, NO. 3, MARCH 2003
production III–V HBTs [2], [3]. The possibility to design the
very high-performance bipolar device on the same chip as high
performance CMOS and with it, the large menu of passive
elements including resistors, capacitors, inductors, varactors,
and low-loss transmission lines opens up numerous possibilities that are not attainable in III–V technologies. For instance,
the high-performance optical networking parts, in addition to
containing the very high-speed functional blocks served well by
the high-speed bipolar transistors, also have a need for de-skew
logic, self-test, and forward error correction. These elements
may be implemented with a low-power solution in a relatively
small amount of CMOS logic, but would be prohibitive to
implement in bipolar transistors. Lower speed analog elements
more favorable to implementation in CMOS, such as op-amps
and quadrature voltage-controlled oscillators (VCOs) are also
possible. Further, the on-chip digital CMOS supports the trend
to integrate more and more of the framer logic onto the same
chip as the high-speed physical layer portions. By having
available the wide range of devices in BiCMOS technologies
on a single chip, system architectures have greater flexibility,
and power hungry and signal degenerating off-chip connections
are avoided. Also important is that complex design system
discontinuities are avoided as is required to design systems
in multiple chips. The capability to combine the high-speed
bipolar with the CMOS will be a strong motivator to implement
in silicon technologies wherever possible.
Obtaining such high performance with a silicon bipolar transistor does not come without breaking some perceived barriers.
For instance, the well-known “Johnson limit” predicts that silicon (also SiGe) bipolar transistors will be limited to an
product of 200 GHz V [3]. This of course predicts a
of below 1 V for 200-GHz transistors, yet this is now
products exceeding
being routinely broken with
values for over
350 GHz-V [1], [5], [6]. Still, typical
transistors are less than 2 V, and this has caused
100 GHz
some concerns to designers whose load lines extend beyond this
value of collector–emitter voltage.
Higher current densities compared to prior generations of silicon bipolar transistor also have caused some concern. Although
in each transistor generation transistor dimensions shrink while
the current density increases (see, for example, Fig. 1), it has
been found that transistor parameter degradation can be exacerbated by this current density.
In this paper, we first describe SiGe HBT transistor design,
covering critical attributes essential to producing greater than
and
. In the process, we compare to
200 GHz in both
both other published SiGe HBT results and InP HBT results. We
then discuss such application-related issues as just described,
addressing the viability of over 200-GHz transistors in applica, and retions swinging collector–emitter voltages over
liable operation of these transistors at high current density. We
then describe some ring-oscillator results demonstrating record
stage delays.
II. TRANSISTOR DESIGN
In high-speed transistor design, it is beneficial to decouple
optimization from the task of base resistance and
the task of
Fig. 2. Scaling the SiGe HBT involves reducing the base width, increasing the
Ge ramp, and increasing the collector concentration. Simultaneous structural
are also
changes to improve in R and R and minimize increase in C
required for highest performance.
collector–base capacitance reduction. The former requires attention to the vertical profile, while the latter is a task of creating a transistor structure that minimizes the parasitic elements
optiand at the same time, does not negatively influence the
mization covered in the former task. The benefit not only holds
in a method to describe the elements of design, but also in the
transistor development procedures, as was done with development of the transistor reported here (see sequential reports [7]
and [1]). In this section, we compare our 200-GHz transistor
design to our 120-GHz transistor design, to illustrate the significant modifications to prior state-of-the-art.
A. Vertical Profile Design
improvement is achieved through improvement in each
in the well-established
of the delay components, related to
relation
where
itance,
and
are emitter–base and base-collector capacand
are collector and emitter resistance,
and
are neutral base and base-collector space charge layer
width, respectively, is Boltzmann’s constant, is temperais
ture, is the unit electron charge, is field factor, and
the electron saturation velocity.
(base transit time) and
The transit time elements
(base-collector space-charge layer transit time) are reduced
through vertical dimension minimization and Ge profile optimization. The base film is deposited through a low-temperature
UHVCVD, incorporating a thin boron layer with a graded
Ge concentration. As shown in Fig. 2, a thinner as-deposited
FREEMAN et al.: TRANSISTOR DESIGN AND APPLICATION CONSIDERATIONS FOR
(a)
200-GHz SiGe HBTs
647
(b)
3
Fig. 3. f versus (a) J (b) and f versus linear current (=J
W ) for
0.12
2.5 and 0.2
6.4 m transistors. Note that the plot on the right is
relevant to the metal system, which provides similar current per unit transistor
length for both transistors.
2
2
boron results in a smaller base width after processing and base
broadening by diffusion, and the graded Ge profile results
in an accelerating field for electrons across the base, both of
component of delay.
which have a strong effect on the
Carbon is incorporated within the base layer during growth in
order to minimize undesirable boron widening in subsequent
high-temperature process steps [8], [9]. Improvement of
involves reducing the base-collector space-charge layer width,
and this is accomplished by increasing the collector doping
transistor, we
concentration. Compared to our 120-GHz
term by 37% to
have been able to reduce the
0.48 ps.
Increasing the collector concentration has additional performance enhancing benefits. First, it enables the transistor to operate at higher currents (pushing out the Kirk effect) and thus
term of the above relaminimizing the
tion. Second, together with reduction in emitter resistance and
minimization of collector epitaxy thickness, it reduces the access resistance portion of the delay relation
.
The key difficulty in increasing the collector dose is that it
component as well. Significant improveincreases the
ments incorporating increased collector dose thus must go
hand in hand with structural or layout modifications, such
is not increased substantially with increased dose.
that
Layout improvements, such as reduced emitter and collector
implant (pedestal) dimensions, have limited effect due to the
perimeter components of capacitance. Structural and process
are
modifications that reduce the extrinsic portions of
performance in SiGe
essential to achieving the highest
HBTs, and these will be the topic of the next section.
Self-heating and metal interconnect current capability are key
considerations in creating a reliable transistor and dictate that,
with increased current density in the transistor, a more narrow
emitter stripe width is required. Shown in Fig. 3 is a comparison
scale and
of the two technologies on both a current density
on a current scale, defined as the product of the current density
and the emitter stripe width
. The
value approx-
Fig. 4. Estimated ECL delay components for the 120- and 210-GHz
technologies. Leftmost bars are the portion from transistor transit times .
The remaining four are the R C delay components, where C is the weighted
capacitance charged through the associated resistance.
3
imates the current through a unit width of metal, corresponding
to a unit length of the emitter. Design guidelines specify a maximum current per unit metal width for reliable operation. Maintaining a similar value between generations is required to maintain reliable operation of the metal system, and thus the more
scale.
important comparison is on the
B. Transistor Structure Improvements
Delay in a circuit is not well predicted by the transistor paalone and is a function of the device parasitic acrameter
cess resistances and parasitic capacitances as well. We have utilized the method described in [10] and [11] to estimate the delay
components in an ECL circuit as shown graphically in Fig. 4.
delays,
This shows the transit-time related delay and the
is the load resistor,
is the base resistance,
where
the collector resistance, and
the emitter resistance.
,
,
, and
, are the total weighted capacitances to
be charged through the load, base, collector, and emitter resistances according to this technique. The comparison between
the 120-GHz technology and the 210-GHz technology in this
chart is illustrative of several important observations. First, the
with the asso120-GHz technology is largely limited by
ciated capacitances. These components have been greatly improved moving to the 210-GHz device, yet remain as significant delay limiters. Note that, even though the base resistance is
a strong limiting resistance in the device, it does not influence
figure of merit. Improvement in this key parameter will
the
be discussed shortly.
figure of merit attempts to be
As is well known, the
more relevant to capture such RC delays as highlighted in the
above analysis. Reduction in base resistance, without sacrificing
or , is the critical challenge. It is interesting to compare
generations of technology in these key parameters. With genervalues,
ations of technology represented by their respective
and
versus the technology
we plot in Fig. 5 the
648
Fig. 5. Unit length C
and R versus f . C
increases only slightly
despite the higher dose required between generations. R decreases between
generations to improve circuit speed. Also note that the R and C
for the
210-GHz SiGe transistor are comparable to the best reported InP (nontransferred
substrate) devices [1], [3], [12]–[17].
for several published technologies. Note that in this plot,
is decreased between generations, and
remains relatively
constant despite the increased collector dose required to achieve
values.
the higher
Also shown in this plot are InP HBT published results. Contrasting the mesa structures of InP technology, where the base
sheet resistance is low, yet limited extrinsic base resistance lowbetween
ering options are available, results in similar total
is lower in InP
the material systems. Similarly, unit area
technologies, yet it is uniform over the total area under the device. This compares to silicon bipolar devices where the capacitance is dominated by the patterned pedestal implant under a
small portion of the total device. The lowest InP points in the
chart are not apparent to be production devices, due to the substantial shrink and dependence on small features of wet etches
to achieve such low numbers. Structure flexibility in silicon processing appears to have matched or overcome the material advantages of InP transistors.
For the 210-GHz device shown here, these low values for
and
result in an
value of 285 GHz (extracted
from Mason’s unilateral power gain). Such values are achieved
through a new self-aligned raised extrinsic base structure, which
has substantial benefits beyond other common structures. One
common structure, where the extrinsic base is implanted into the
SiGe epitaxy layer, is represented in the 120-GHz device used
for reference in this paper. This is also utilized in technologies
from other companies [5]. Another common structure is characterized by selective epitaxy following an opening in a low-resistance extrinsic base polysilicon film [6], [14], [12]. Both these
structures appear to have limiting issues as will be described
here.
We have found that, in the implanted extrinsic base structure,
and
are both limited by the implant and that
cannot
easily be decoupled from the extrinsic base construction. The
implanted extrinsic base structure is shown in Fig. 6(a). This
figure depicts two key issues found in this structure. First, the
depth of the implant is necessarily deeper than the intrinsic region, and thus the collector pedestal will intersect this and col-
IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 50, NO. 3, MARCH 2003
Fig. 6. (a) Implanted extrinsic base device structure. (b) Raised extrinsic base
structure.
lector–base capacitance is increased. A shallower implant or
for lower
.A
further lateral distance trades off higher
second factor is the implant damage and the effect this has on
the intrinsic region of the device. Implant damage creates silicon interstitials which migrate to the base region of the device
and create excess diffusion, increasing base transit time, and reducing . Thus, the implanted extrinsic base has a wide influence on a variety of parameters and thus limits are found to
and
.
achievable
Similarly, the selective epitaxy approach has its limitations.
As described in [18] and [19], the growth and link-up of the selective film on the underside of the extrinsic base polysilicon
presents issues. First, growth defects such as twins and silicon
dislocations are found in close proximity to the device and the
collector–base junction, and this causes emitter–base and collector–base leakage issues. Second, the thickness of the epitaxy intrinsic layer put down prior to boron layer growth is constrained, since this also grows on the underside of the extrinsic
base polysilicon. A thicker film will make link-up between the
extrinsic polysilicon and the selective epitaxy film more difficult. Distance to the emitter and extra thermal processing may
be required to minimize the effects. Thus, the base resistance
optimization.
minimization is not decoupled from the
The raised extrinsic base structure utilized on the 210-GHz
device described here contains a raised base that is deposited
after the SiGe epitaxy film, and is self-aligned to the emitter.
This is illustrated in Fig. 6(b). Because there is no implant into
resulting from
the SiGe epitaxy film, there is little impact on
proximity of the extrinsic base to the emitter. In addition, because the extrinsic base diffuses into the SiGe film only a small
amount, the base side of collector–base junction is the same as
is minimized. Comthe intrinsic base dopant. In this way,
pared to the selective epitaxy processes, the SiGe epitaxy film
is deposited across the entire wafer, and most importantly, the
active device and junctions are contained in a film that is influenced strictly by growth on single crystal. The linkage between
the raised extrinsic base and the intrinsic base film occurs over
a larger distance. These various factors are significant to make
the next step in silicon bipolar improvements.
FREEMAN et al.: TRANSISTOR DESIGN AND APPLICATION CONSIDERATIONS FOR
Fig. 7. (a) Gummel and (b) output characteristics for a 0.12 m
207-GHz f device.
2 2.5 m
200-GHz SiGe HBTs
649
Fig. 8. (a) f and f
versus collector current and Mason’s unilateral gain
(U) and (b) maximum stable and maximum available gain (MSG/MAG) versus
frequency at peak bias.
C. Device Results
The devices reported here have pinched-base sheet resistance
of 1.7 V,
of 2.5 k /sq, a peak dc current gain of 400,
of 5.5 V. Shown in Fig. 7 are dc characteristics for
and
an NPN with 0.12 2.5 m emitter area. The Gummel characteristic shows a collector current ideality of 1.02 and a base
current ideality of 1.15. The negative slope in the output characteristics reflects self-heating in the device, which is discussed
in more detail in a later section and in [20].
On-wafer high-frequency measurements up to 110 GHz were
performed. Measured pad parasitics were de-embedded at every
frequency point using on-chip open-short calibration. Fig. 8
and
(both U and MAG) for
shows room temperature
of 1 V.
a device with a 0.12 2.5 m size emitter at a
in excess of 200 GHz is seen for a current density range
of 8.3–16.5 mA/ m (linear current of 1.0–1.98 mA/ m),
in excess of 270 GHz is achieved
peaking at 207 GHz.
from the 20 dB/dec extrapolation of Mason’s unilateral power
gain (U) at 40 GHz for the same current range with a peak value
extrapolation from maximum
of 285 GHz. Alternatively,
available gain (MAG) in the 50–80-GHz range of frequencies
values of 194 GHz. The actual values
results in peak
for power gain at various discrete frequencies of interest are
to understand
perhaps more practical than extrapolated
the real-world performance potential of a device. Fig. 8 also
illustrates both U and MAG versus frequency for the 0.12
2.5 m device, revealing values for U and MAG at 40 GHz of
17.0 and 15.9 dB, respectively. These values advance the SiGe
HBT state of the art by 4-5 dB compared to previously reported
figures in the range of 160–170 GHz [14], [5]. The gain
data exhibit a 20 dB/dec slope in U between 20–60 GHz and
of 50
a 23 dB/dec slope in MAG between 50–80 GHz.
and
of 4.2 fF is obtained from measured -parameters
on this same device.
Small device performance is key to achieving large transistor
counts on a single chip with reasonable overall power dissipation. Fig. 9 shows ac characteristics for a small device (0.12
0.5 m at 1 V
. A peak
of 180 GHz is obtained
Fig. 9. (a) f and f
versus collector current for a small device (0.12 m
0.5 m), with (b) measured h power gain versus frequency at peak bias.
2
at 800- A current. For applications requiring bias with equivtechnology (e.g., 40 Gb/s netalent gain as the 120-GHz
at a much lower curworking), this device provides such an
rent of 175 A, which is a 78% decrease from the prior gencharacteristics are well beeration. As shown in Fig. 9,
haved at 20 dB/dec roll-off, yet measurements on such small
devices are problematic in obtaining accurate U, shown for refbias. From these measurements
erence in Fig. 9(b) at peak
and
for
and comparison of the product of extracted
both large and small devices, we estimate this smallest device to
values equal to or greater than the longer devices.
have
Some transistor designs, utilizing for instance low base
dopant concentration or nonmanufacturable dimensions or
overlay requirements, may achieve high performance yet not
have sufficient robustness to achieve the same performance
over normal process variations. The design should be robust
enough to ensure device performance in the limits when process
650
IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 50, NO. 3, MARCH 2003
Fig. 11. f and f
Fig. 10.
R
f
and f
= 3500 /sq.
versus pinched-base sheet resistance R
.
across a wafer and with different emitter widths.
parameters drift, resulting in unintentional modifications to
dopant diffusion, layer thickness, or lithography dimensions.
The design described thus far utilizes high base dopant concentration and, as will be shown, such a doping level ensures
) and
minimal variation in neutral base width (and hence
avoids punch-through concerns even for relatively high values
. Additionally, the design takes into account the tolerof
ances of the emitter–base junction depth and collector–base depletion width, resulting in a well-controlled collector current
to the first order is affected by
and . The tolerance of
and
distributions. Process variations in
are acthe
counted for by understanding the tolerances of its intrinsic and
extrinsic components. The intrinsic component depends on the
intrinsic base doping and the definition of the emitter dimension.
The emitter dimension and its separation from the extrinsic base
are defined by standard CMOS processes and thus well-controlled dimension variations are ensured. In addition, because
is
of the high base doping, the intrinsic portion of the total
is further insensitive to the emitter width.
small and thus
Formation of the raised extrinsic base with a shallow junction
depth and minimal TED results in a tight distribution of the
and
. Furthermore, the
extrinsic component of both
raised extrinsic base scheme minimizes the effects of variable
and unwanted slower perimeter component of device operation
by minimizing the perimeter base diffusion effects and controlling the counterdoping of the emitter at the perimeter.
and
dependence
Shown in Figs. 10 and 11 are the
, reon emitter width and pinched-base sheet resistance
spectively, for a set of wafers run to explore these dependencies.
is slightly different from that reported
The nominal value of
above due to slight process differences. Fig. 10 shows relatively
and
, where the
variation is
tight distribution in
largely due to uncertainties in the unilateral gain due to meawith
surement variability. Approximately 5% change in
33% larger emitters is observed, and this is due to an increase
from the larger area device. Extracted
appears to be
in
similar between emitter widths.
Further, we show process tolerance as a function of nominal
pinched-base sheet resistance design points through simulation.
values with high pinched-base sheet
One can obtain high
Fig. 12. Influence on I and R
with variations in base and emitter
concentrations for device designs of nominal 4 and 20 K
/sq.
resistance (low base doping and narrow base widths), yet device sensitivity to process variations is sacrificed. Numerical
simulations [21] were employed to compare device sensitivities
with nominal pinched-base resistance of 4 and 20 K /sq. For
values, the base and
boron doses corresponding to such
emitter dopant concentrations were varied 20% and 20%,
representing expected process variations. Such variations result
values and in device parameter variations as
in different
well. Shown in Fig. 12 is a plot of the collector current for
these individual process variations as a function of the
value. Simultaneous variations in process parameters will result
in greater variations in transistor properties. From this plot, it
is observed that sensitivities to process variation become subdesign point. Greater base
stantially greater at the higher
design point rewidth sensitivity exhibited in the higher
as shown, but also in , early
sults in variations not only in
, making the lower
design point prevoltage, and
ferred for a manufacturable process.
III. APPLICATION ISSUES
The high-performance SiGe HBT described in this paper has
value and a higher current density as coma reduced
pared to prior generations of slower devices. The devices devalues of approximately half
scribed here demonstrate
FREEMAN et al.: TRANSISTOR DESIGN AND APPLICATION CONSIDERATIONS FOR
that of the 50-GHz
devices of prior generation and seven
times the current density of the same generation. This section
addresses concerns relating to such trends necessary to achieve
the higher performance, and shows that designs utilizing higher
and higher current denperformance devices with lower
sity are possible and that the device degradation under such conditions is well within acceptable specifications.
200-GHz SiGe HBTs
651
TABLE I
ESTIMATED JUNCTION TEMPERATURE RISE AT PEAK f BIAS CURRENT AND
V = 1:5 V FOR VARIOUS 150–200-GHz HBT TECHNOLOGIES, WITH
DEVICE SIZES CHOSEN FOR THE SAME 6-mA PEAK I [20], [25]
A. Current Density
High current density causes concern in two areas: 1) device
parameter degradation with long-term operation and 2) HBT device self-heating and impact on degradation and long-term reliability. While these effects are related, since the latter influences
the former, the high current (without the heat) directly influences device degradation and may be viewed separately.
Relative to current-induced device degradation, many concerns appear to come from the experience with III–V devices
where high current density has historically caused severe device
degradation and sudden collapse of gain. The observed mechanism has been the migration of dopant atoms in the crystal lattice of the base film, with movement of the dopant atom beyond
the heterojunction causing catastrophic effects [22]. One can
easily understand that this effect does not occur in silicon and
silicon-germanium systems since the bonding energy of silicon
systems is so much greater than III–V systems. This is observed
in the difference in diffusion constants of dopants and in the
practical fact that silicon systems must achieve 900 C temperatures for significant diffusion to occur, compared to maximum
allowable temperatures of 400 C–500 C for III–V systems.
Degradation mechanisms observed for silicon systems in forward bias are quite different, and appear to relate to hydrogen
trapping and detrapping in the polysilicon to single crystal interfacial region as a function of current density. As reported in [23],
[24], this effect results principally in an ideal base current shift.
Effects on collector current appear to be observed principally
at higher currents where emitter resistance shifts are observed.
Degradation in devices stressed at peak current densities of 8
mA/ m has been detailed in [24], with junction temperatures
tested up to 235 C, and extrapolations to over 10 h with acceptable degradation. With careful attention to engineering the
polysilicon/single-crystal interface, the degradation can be minimized as current densities are increased. Similar device stress
results are currently under investigation for the 200-GHz transistors described here.
High current densities are also responsible for heating of the
device. The heating of the device will translate to the metal
system, where a higher temperature will cause earlier failure of
the metal interconnect above the device.
Table I compares self-heating approximations across different 150–200-GHz technologies. Namely, the SiGe HBT
device described here is compared to the reported self-heating
characteristics of InP HBTs [25]. Since thermal resistance in
InP devices is a strong function of the collector design, with
InGaAs collectors or subcollectors inhibiting heat dissipation
compared to InP collectors, both types of devices are shown.
In this comparison, we normalize to an operating current for a
mA, since normalized thermal dissipation
transistor, at
2
Fig. 13. Comparison of temperature rise for the 0.12 2.5 m SiGe HBT
transistor and 1
3 m InGaAs collector InP HBT transistors presented
in [25]. Of the two InP transistors, the higher f values result from reduced
parasitics from improved lithography overlay and smaller base mesa structure.
2
- m varies with device size. The InP thermal resistance values are comparable to those reported in [26]–[28].
Note that although the typical operation current densities
between material systems are very different to achieve peak
, the thermal resistance of an InP device with the InGaAs
collector is very high and results in very large self-heating. In
Fig. 13, the results for this type of collector design are shown
from [25] compared to the 200-GHz SiGe HBT self-heating
characteristics. The all-InP collector design results in compavalues from
rable self-heating at peak to the SiGe HBT, yet
production transistors with InP collector are approximately
20% lower than in the SiGe HBT reported in this paper.
For the SiGe transistor, the principle engineering solution to
the increasing current density is to narrow the emitter stripe
width inversely with the current density, resulting in a fixed current along the length of the device from one generation to the
next. Because of the narrowed width, the device lengthens for
a given emitter area or device peak current, and the thermal resistance is reduced. Still, as shown in Fig. 14, the thermal resistance does not scale directly with the stripe width and as a
result, a device with higher current density will become hotter
operating at peak current compared to a prior generation [20].
Alternate solutions are also feasible. An increase in performance is one alternative, such as shown in Fig. 3, where the
652
IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 50, NO. 3, MARCH 2003
Fig. 14. Measured (symbols) and calculated (lines) thermal resistance for
various device dimensions. From [20].
improved device operates at lower current densities for comparable performance. This provides the designer the flexibility to
use a larger device with lower current density, (and same total
current), achieving the same high-frequency gain.
Further, careful thermal considerations in circuit design can
ensure reliable operation. For this, the designer knows that the
backside of the chip is maintained at a certain temperature, and
a temperature rise may be determined for each device in the circuit. Utilizing this method and typical backside temperatures
of 85 C, one finds that most devices are well under the typically assumed junction temperature of 125 C, corresponding
to a 40 C temperature rise due to self-heating and neighboring
device effects. Those devices that operate at high bias and peak
current (or are tightly packed) require extra attention in design,
by spreading the heat into a larger footprint in the device design
or through special attention to the metal interconnect (such as
strapping metal levels together).
Guidance may be provided to the designers from the transistor model, incorporating circuit operating conditions and
thermal resistance dependencies on geometry. The self-heating
typically incorporated in the model may be used to highlight
devices which have become the hottest.
Alternatively, graphical methods may be utilized. One graphversus
, is instructive to
ical method, applied to a plot of
define what constraints exist on where a device may be biased.
and high region of the plot is the
On such a plot, the high
region of high self-heating. The typical electromigration guideline of a maximum current as a function of temperature can take
into account the self-heating of the device (which is a function
of bias) and a solution to the electromigration and self-heating
equations may be plotted to define the limit. Such a plot is defined in Fig. 15 for the 200-GHz device design and a copper
metal system with parts per trillion failure rates, as well as moderate packing densities. One can observe that, while not all desired biasing options are possible with the metal configuration
used to define this set of constraints, the device can nonetheless
be operated such that peak performance and reliable operation
may be obtained. Note that the limit does not result from the intrinsic device characteristic, but is related to an extrinsic factor,
Fig. 15. Bias constraints for the metal system on a simple one-stripe device
layout, taking into account self-heating of the device as a function of bias
condition and the dependence of the metal system reliability on this device
temperature. All data is for 0.12-m-wide emitters, with the length shown in
the legend. The load line is for the 0.12 2 m device in the ring-oscillator
circuit of Fig. 16.
2
which is the characteristic of the metal interconnect. This implies that the limit can be relaxed with layout modifications.
B. Breakdown
Even though the Johnson limit of 200 GHz V is now being exvalues of these deceeded by today’s 200-GHz devices,
vices remain less than 2 V. Designers thus must face the prospect
. It has been
of designing load lines that extend beyond
is not a fundamental limit
pointed out by others that
for allowed bias voltage, but that it simply marks the value of
where the base current reverses because of avalanche multiplication in the collector–base space-charge region. Instead,
have been explored for the
practical limits larger than
device, and realistic reliability issues have been defined. These
are covered in detail in [29].
One practical issue is in the effect of the negative base current
on the device bias. Device output conductance is increased because of the avalanche current passing through the internal and
external base resistance components, and further turning on the
base-emitter junction. This effect may be managed well through
careful device modeling, such that the designer is made aware
of the effects of the base current.
Another issue related to the base current reversal is the device
“pinch-in,” where the potential drop resulting from avalanche
current flowing out of the device through the internal base resistance, eventually drives the large majority of collector current
through the center of the device, resulting in hot-spots during
operation. This is apparent at voltages above 3.5 V for the devices described here.
Power amplifier designs and modulator drivers routinely
limits to achieve high voltage swings
break the
FREEMAN et al.: TRANSISTOR DESIGN AND APPLICATION CONSIDERATIONS FOR
Fig. 16. Stage delay versus collector current (tail current) for ring oscillators
fabricated with 0.12- and 0.16-m emitter width (enw) NPN and different load
resistors (R ). The inset shows the schematic of a single ECL gate.
approaching the device
. Power amplifiers utilizing
GHz devices with
V and
roughly
achieve voltage swings up to 20 V at a 2-GHz operating
frequency [30]. Modulator drivers operating to 40 Gb/s have
also been reported, and these demonstrate 2.5–3.5 V peak–peak
and 5 V voltage
voltage swings, requiring between
and
,
swings on devices with 1.8 and 6.5 V
respectively [31], [32].
IV. CIRCUIT RESULTS
We describe here record ring-oscillator results of a ring-os-GHz device described
cillator fabricated with the
earlier. The ring oscillator implemented here is based on a standard ECL topology. A single ECL stage consists of seven NPN
transistors, all of which are 0.12 m 2 m in size (see the
inset of Fig. 16). The design consists of 14 stages and is fully
differential. The tail current (collector current) in the differential pair is set up through a current mirror that generates current
proportional to an input voltage. Further, the bias conditions are
such that device self-heating is kept below the limits of the metal
system (Fig. 15). It is estimated that 1 mA is consumed in the
emitter follower of the ECL stage at a 3.6-V nominal supply
voltage. ECL stage-to-stage wiring was done in the first level of
metal.
Fig. 16 compares room-temperature (25 C) stage delay
curves for ring oscillators with emitter width (enw) of 0.12 m
combinations. For
and 0.16 m with various load resistor
a 3.6-V supply, minimum delay of 4.3 ps is obtained for rings
(as designed) at
designed with 0.12- m emitters and 105
a collector current of 2.2 mA. Increasing power supply to 4.0
V reduces the delay further to 4.2 ps. As expected, increasing
the load resistor values leads to an increase in delay times. All
other aspects of the design remaining the same, increasing the
emitter width to 0.16 m leads to a 5% increase in delay times,
reduction in larger device
which is comparable to the
200-GHz SiGe HBTs
653
Fig. 17. (a) ECL stage delay comparison for ring oscillators fabricated in 90SiGe technologies. (b) Bipolar ring-oscillator stage delay
and 285-GHz f
as a function of f
(unilateral gain) from this work and recently reported
data in literature. References A = [34], B = [35], C = [36].
sizes. From -parameter measurements, a
increase of
10% results from increasing the emitter width to 0.16 m, and
and the rise
this contributes to the observed reduction in
in delay.
ECL ring oscillators were also fabricated in a previous
and 90-GHz
(unilateral gain) technology
90-GHz
4 m with
[33]. This design uses NPN devices sized 0.2
and was biased from a 3.6-V power supply as
100
well. The prior generation NPNs have 2 larger transit time,
2 higher base resistance, 2 higher emitter resistance, 2
higher collector resistance, and 35% lower
compared
to the 200-GHz NPN devices. Fig. 17(a) compares the ECL
delays in the two technologies. Similar to the delay component
breakdown shown in Fig. 4, we find that a large portion of
the 50% delay reduction is a result of the base resistance
improvement. The strong dependence of ring-oscillator delays
on NPN parasitics is shown in Fig. 17(b), where stage delay
(from unilateral gain) for this work as
is plotted against
well as recent ECL and CML gate delay data reported in the
literature.
V. CONCLUSION
We have shown that 200-GHz devices can be achieved in
SiGe systems and discussed the relevant aspects of device design that are important to achieve high performance, manufacturable devices, and reliable systems. Compared to InP devices,
and
and, importantly, opthese devices have higher
erate at substantially lower temperature rise.
In transistor design, minimization of parasitic elements improviding the greatest leverage.
proves performance, with
Common silicon processing techniques are employed to obtain
minimized extrinsic base resistance, and minimal increase in
collector–base capacitance. Due to such techniques, we find that
silicon transistors can surpass III–V device performance despite
654
IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 50, NO. 3, MARCH 2003
the well-known III–V material system benefits, and bring with
it the integration benefits well known to silicon technology.
[21] J. Johnson, A. Stricker, A. Joseph, and J. Slinkman, “A technology simulation methodology for AC-performance optimization if SiGe HBTs,”
in IEDM Tech. Dig., 2001, pp. 489–492.
[22] M. Hafizi and M. J. Delaney, “Reliability of InP-based HBT’s and
HEMT’s: Experiments, failure mechanisms, and statistics,” in Proc.
Int. Conf. InP and Related Materials, 1994, pp. 299–302.
[23] M. S. Carroll, A. Neugroschel, and C.-T. Sah, “Degradation of silicon
bipolar junction transistors at high forward current densities,” IEEE
Trans. Electron Devices, vol. 44, pp. 110–117, Jan. 1997.
[24] J.-S. Rieh, K. Watson, F. Guarin, Z. Yang, P.-C. Wang, A. Joseph, G.
Freeman, and S. Subbanna, “Wafer level forward current reliability
analysis of 120GHz production SiGe HBT’s under accelerated current
stress,” in Proc. IEEE Int. Reliability Physics Symp., Apr. 2002, pp.
184–188.
[25] S. Thomas, III, J. A. Foschaar, C. H. Fields, M. M. Madhav, M.
Sokolich, R. D. Rajavel, and B. Shi, “Effects of device design on
InP-based HBT thermal resistance,” IEEE Trans. Device Mater. Reliab.,
vol. 1, pp. 185–189, 2001.
[26] W. Liu, H. F. Chau, and E. Beam, III, “Thermal properties and thermal
instabilities of InP-Based heterojunction bipolar transistors,” IEEE
Trans. Electron Devices, vol. 43, pp. 388–395, Mar. 1996.
[27] S. R. Bahl, N. Moll, V. M. Robbins, H. C. Kuo, B. G. Moser, and G.
E. Stillman, “Be diffusion in InGaAs/InP heterojunction bipolar transistors,” IEEE Electron Device Lett., vol. 21, pp. 332–334, July 2000.
[28] M. Ray, D. Hill, O. Hartin, K. Johnson, and P. Li, “Thermal impact of InGaAs on InP based HBTs,” in Proc. GaAs IC Conf., 2001, pp. 261–264.
[29] M. Rickelt, H. M. Rein, and E. Rose, “Influence of impact-ionization-induced instabilities on the maximum usable output voltage of Si-Bipolar
transistors,” IEEE Trans. Electron Devices, vol. 48, pp. 774–783, Apr.
2001.
[30] D. R. Greenberg, M. Rivier, P. Girard, E. Bergeault, J. Moniz,
D. Ahlgren, G. Freeman, S. Subbanna, S. J. Jeng, K. Stein, D.
Nguyen-Ngoc, K. Schonenberg, J. Malinowski, D. Colavito, D. L.
Harame, and B. Meyerson, “Large-signal performance of high-BV/sub
CEO/ graded epi-base SiGe HBT’s at wireless frequencies,” in IEDM
Tech. Dig, 1997, pp. 709–802.
[31] R. Schmid, T. F. Meister, M. Rest, and H.-M. Rein, “40 Gbit/s EAM
driver IC in SiGe bipolar technology,” Electron. Lett., vol. 34, no. 11,
pp. 1095–1097, May 1998.
[32] G. Freeman, M. Meghelli, Y. Kwark, S. Zier, A. Rylyakov, M. A. Sorna,
T. Tanji, O. Schreiber, K. Walter, J.-S. Rieh, B. Jagannathan, A. Joseph,
and S. Subbanna, “40 Gbit/s circuits built from a 120 GHz f SiGe
technology,” IEEE J. Solid State Circuits, vol. 37, pp. 1106–1114, Sept.,
2002.
[33] G. Freeman, D. Ahlgren, D. R. Greenberg, R. Groves, F. Huang, G.
Hugo, B. Jagannathan, S. J. Jeng, J. Johnson, K. Schonenberg, K.
Stein, R. Volant, and S. Subbanna, “A 0.18 m 90 GHz f SiGe HBT
BiCMOS, ASIC-compatible, copper interconnect technology for RF
and microwave applications,” in IEDM Tech. Dig., 1999, pp. 569–572.
[34] J. Bock, H. Schafer, H. Knapp, D. Zoschg, K. Aufinger, M. Wurzer, S.
Boguth, R. Stengl, R. Schreiter, and T. F. Meister, “High-speed SiGe : C
bipolar technology,” in IEDM Tech. Dig., 2001, pp. 344–347.
[35] E. Ohue, R. Hayarni, K. Oda, H. Shimarnoto, and K. Washio, “5.3-ps
ECL and 71-GHz static frequency divider in self-aligned SEG SiGe
HBT,” in Proc. BCTM, 2001, pp. 26–29.
[36] M. Sokolich, A. R. Kramer, Y. K. Boegeman, and R. R. Martinez,
“Demonstration of sub-5 ps CML ring oscillator gate delay with
reduced parasitic AlInAs/InGaAs HBT,” IEEE Electron Device Lett.,
vol. 22, pp. 309–311, 2001.
REFERENCES
[1] B. Jagannathan, M. Khater, F. Pagette, J.-S. Rieh, D. Angell, H. Chen,
J. Florkey, F. Golan, D. R. Greenberg, R. Groves, S. J. Jeng, J. Johnson,
E. Mengistu, K. T. Schonenberg, C. M. Schnabel, P. Smith, A. Stricker,
D. Ahlgren, G. Freeman, K. Stein, and S. Subbanna, “Self-aligned SiGe
and 207GHz f in a manufacNPN transistors with 285GHz f
turable technology,” IEEE Electron Device Lett., vol. 23, pp. 258–260,
May 2002.
[2] W. E. Stanchina, M. Sokolich, and K. R. Elliott, “Process and integration
technologies for InP ICs,” in Int. Conf. on Indium Phosphide and Related
Materials, 2001, pp. 489–492.
[3] D. Sawdai, E. Kaneshiro, A. Gutierrez-Aitken, P. C. Grossman, K. Sato,
W. Kim, G. Leslie, J. Eldredge, T. Block, P. Chin, L. Tran, A. K. Oki,
and D. C. Streit, “High performance, high yield InP DHBT production
process for 40 Gb/s applications,” in Proc. Int. Conf. Indium Phosphide
and Related Materials, 2001, pp. 493–496.
[4] E. O. Johnson, “Physical limitations on frequency and power parameters
of transistors,” RCA Rev., vol. 26, pp. 163–177, June 1965.
[5] M. Racanelli, K. Schuegraf, A. Kalburge, A. Kar-Roy, B. Shen, C. Hu,
D. Chapek, D. Howard, D. Quon, F. Wang, G. U’ren, L. Lao, H. Tu, J.
Zheng, J. Zhang, K. Bel, K. Yin, P. Joshi, S. Akhtar, S. Vo, T. Lee, W.
Shi, and P. Kempf, “Ultra high speed SiGe NPN for advanced BiCMOS
technology,” in IEDM Tech. Dig., 2001, pp. 336–338.
[6] K. Oda, E. Ohue, I. Suzumura, R. Hayami, A. Kodama, H. Shimamoto, and K. Washio, “Self-aligned selective-epitaxial-growth
Ge C HBT technology featuring 170-GHz f
,” in
Si
IEDM Tech. Dig., 2001, pp. 332–335.
[7] S. J. Jeng, B. Jagannathan, J.-S. Rieh, J. Johnson, K. T. Schonenberg, D.
Greenberg, A. Stricker, H. Chen, M. Khater, D. Ahlgren, G. Freeman,
K. Stein, and S. Subbanna, “A 210 GHz f SiGe heterojunction bipolar
transistor with a nonself-aligned (NSA) structure,” IEEE Electron Device Lett., vol. 22, pp. 542–544, 2001.
[8] L. D. Lanzerotti, J. Sturm, E. Stach, R. Hull, T. Buyuklimanli, and C.
Magee, “Suppression of boron outdiffusion in SiGe HBT’s by carbon
incorporation,” in IEDM Tech. Dig, 1996, pp. 249–252.
[9] H. Rucker, B. Heinemann, D. Bolze, D. Knoll, D. Kruger, R. Kurps, H.
J. Osten, P. Schley, B. Tillack, and P. Zaumseil, “Dopant diffusion in
C-doped Si and SiGe: Physical model and experimental verification,” in
IEDM Tech. Dig., 1999, pp. 345–348.
[10] W. Fang, “Accurate analytical delay expressions for ECL and CML circuits and their applications to optimizing high-speed bipolar circuits,”
IEEE J. Solid-State Circuits, vol. 25, pp. 572–583, 1990.
[11] M. Rodwell, M. Urteaga, Y. Bester, T. Mathew, P. Krishnan, D. Scott,
S. Jagannathan, D. Mensa, J. Guthrie, R. Pullela, Q. Lee, B. Agarwal,
U. Bhattacharya, and S. Long, “Scaling to InGaAs/InAlAs HBT’s for
high speed mixed signal and mm-wave ICs,” Int. J. High Speed Electron.
Syst., vol. 11, no. 1, pp. 159–215, 2001.
[12] J. Bock, H. Schafer, H. Knapp, D. Zoschg, K. Aufinger, M. Wurzer, S.
Boguth, R. Stengl, R. Schreiter, and T. F. Meister, “High-speed SiGe : C
bipolar technology,” in IEDM Tech. Dig., 2001, pp. 344–347.
[13] M. Sokolich, C. H. Fields, S. Thomas, III, B. Shi, Y.-K. Boegeman, M.
Montes, R. Martinez, A. R. Kramer, and M. Madhav, “A low power
72.8-GHz static frequency divider in AlInAs/InGaAs HBT technology,”
IEEE J. Solid State Circuits, vol. 36, pp. 1328–1334, Sept. 2001.
[14] K. Washio, M. Kondo, E. Ohue, K. Oda, and R. Hayami, “A 0.2m
self-aligned SiGe HBT featuring 107-GHz f
and 6.7-ps ECL,” in
IEDM Tech. Dig., 1999, pp. 557–560.
[15] T. Oka, K. Hirata, K. Ouchi, H. Uchiyama, T. Taniguchi, K. Mochizuki,
and T. Nakamura, “Advanced performance of small-scaled InGaP/GaAs
over 250 GHz,” in IEDM Tech.
HBT’s with f over 150 GHz and f
Dig., 1998, pp. 653–656.
[16] D. Gloria, S. Gellida, and G. Morin, “A new extraction method of high
frequency noise parameters in the temperature range 55/150deg. for
SiGe HBT in BiCMOS process,” in Proc. ICMTS, 2000, pp. 229–234.
[17] G. Freeman, D. Greenberg, K. Walter, and S. Subbanna, “SiGe HBT
performance improvements from lateral scaling,” in Proc. Eur. SolidState Device Research Conf., 1999.
[18] D. Terpstra and W. B. de Boer, “Anomalous collector base leakage in
selectively grown SiGe base heterojunction bipolar transistors,” in Proc.
29th Eur. Solid State Device Research Conf., 1999, pp. 720–723.
[19] K. Oda, E. Ohue, T. Onai, and K. Washio, “Heterojunction bipolar transistor,” U.S. Patent 5 962 880, Oct. 5, 1999.
[20] J.-S. Rieh, D. Greenberg, B. Jagannathan, G. Freeman, and S. Subbanna,
“Measurement and modeling of thermal resistance of high speed SiGe
heterojunction bipolar transistors,” in Silicon Monolithic Integrated Circuits in RF Systems Dig., 2001, pp. 110–113.
0
Greg Freeman (S’82–M’90–SM’03) received the
B.S.E.E. degree from the University of Delaware,
Newark, in 1984 and the M.S.E.E. and Ph.D. degrees
from Stanford University, Stanford, CA, in 1986 and
1991, respectively. His dissertation dealt with data
analysis for semiconductor process diagnosis.
He has been with IBM’s East Fishkill, NY, facility
since 1991 in the development of advanced semiconductor processes, including DRAMs, CMOS, and,
since 1995, SiGe BiCMOS. His recent achievements
include leading a team to develop the world’s first
210-GHz silicon–germanium heterojunction transistor. Since 2000, he has
been a Senior Engineering Manager in IBM Microelectronics, responsible
for communications technology research and development including device
design, circuit applications, and process integration for SiGe HBTs as well as
RF-CMOS. His past and current research interests are in device diagnostics,
detailed device design, and the application of semiconductor devices and
technologies. He has authored or coauthored over 40 technical publications
and presentations in the fields of electrical characterization and SiGe HBT
BiCMOS technology.
FREEMAN et al.: TRANSISTOR DESIGN AND APPLICATION CONSIDERATIONS FOR
Basanth Jagannathan (M’98) received the B.Tech.
(Hons.) degree from the Indian Institute of Technology, Kharagpur, in 1992 and the Ph.D. degree in
electrical engineering from the State University of
New York at Buffalo in 1997. His doctoral work was
on the growth of amorphous and micro crystalline
silicon for solar cells and thin film transistors.
Since 1997, he has been with IBM Microelectronics communication research and development,
Hopewell Junction, NY, where he now an Advisory
Engineer. He has contributed to the development
of SiGe epitaxy and HBT design in IBM’s SiGe BiCMOS technologies. He
is involved with SiGe technology definition and evaluation for high-speed
communication applications. His current interests are toward understanding
technology design and circuit topologies for implementing wireless and wired
communication chips.
Shwu-Jen Jeng received the B.S. degree in materials
science and engineering from National Tsing-Hua
University, Hsinchu, Taiwan, R.O.C., and the Ph.D.
degree in materials science and engineering from the
University of Illinois, Champaign-Urbana, in 1986.
She was a Post-Doctoral Associate studying
the reactive ion etching damage at the IBM T. J.
Watson Research Center, Yorktown Heights, NY,
from 1986 to 1988. She joined IBM Semiconductor
Development Lab, Hopewell Junction, NY, in 1988
and worked on in situ oxide removal and silicidation
of Si and wide bandgap emitter. In 1994, she joined IBM Microelectronics
Semiconductor Research and Development Center and has been working on
SiGe BiCMOS process and technology development.
Jae-Sung Rieh (S89–M’91) received the B.S. and
M.S. degrees in electronics engineering from Seoul
National University, Seoul, Korea, in 1991 and
1995, respectively, and the Ph.D. degree in electrical
engineering from The University of Michigan, Ann
Arbor, in 1999. His doctoral research involved the
development of SiGe-based electrical and optical
devices and their application to microwave and
optoelectronic integrated circuits.
In 1999, he joined IBM Communications R&D
Center, where he has been involved in the design and
characterization of ultrahigh-speed SiGe HBTs for analog and mixed-signal
applications and the analysis and modeling of the thermal and avalanche
behavior of the SiGe HBTs.
200-GHz SiGe HBTs
655
Andreas D. Stricker was born in Bern, Switzerland,
in February 1965. He received the M.S. degree in applied physics from the University of Bern in 1992 and
the Ph.D. degree from the Swiss Federal Institute of
Technology, Zürich, in 2000. His dissertation was titled “TCAD of ESD protection devices.”
While at the University of Bern, he worked
in the field of LASER interactions with solid
state materials. His fields of interest were on-chip
ESD-protection circuits as well as their development
using process and device simulations. He is now with IBM Microelectronics,
Burlington, VT, where he joined the SiGe design kit development group.
David C. Ahlgren received the B.A. frhtrr from DePauw University, Greencastle, IN, in 1973 and the
Ph.D. degree in chemical physics from The University of Michigan, Ann Arbor, in 1979.
He joined IBM, Hopewell Junction, NY, in 1979,
conducting semiconductor process development.
His early work was in the area of silicon defects
resulting from ion implantation and isolation stress,
as well as process integration issues which lead
to the development and subsequent production of
IBM’s first double polysilicon bipolar technology
in 1983. In 1989, his attention turned to Si/SiGe
HBTs as the next step in the advancement of IBM’s bipolar mainframe
semiconductor technology. His early device studies, process technology work,
and semiconductor production experience has lead him into his role in the
Advanced Semiconductor Technology Center as a Senior Engineer in device
and process development of high-performance Si/SiGe BiCMOS technology
and its introduction into manufacturing. He is currently Senior Engineering
Manager of SiGe Advanced BiCMOS Technology Development. He has
published over 30 technical papers and holds eight patents in semiconductor
device and process technology.
Seshadri Subbanna (S’79–M’83) received the is
B. Tech. degree in electrical engineering from the
Indian Institute of Technology, Bombay, and the
M.S.E.E. and Ph. D. degrees in electrical engineering and materials science from the University of
California at Santa Barbara. His doctoral work was
on optical and electronic properties of antimonide
superlattices and heterostructures, including the first
observation of quadratic piezo-electrooptic effect in
strained superlatties.
From 1989 to 2002, he was with IBM Microelectronics, East Fishkill, NY. Since joining IBM, he has worked on advanced
bipolar and CMOS processes, high-density SRAM development, and BiCMOS
technology. From 1996 to 2002, he was manager of advanced silicon-germanium (SiGe) heterojunction bipolar transistor (HBT) BiCMOS development at
the IBM Microelectronics Semiconductor Research and Development Center.
This group was the first to achieve over 200-GHz high-frequency performance
in a silicon-based technology, and his group is recognized world-wide as a
leader in development of high-speed networking technology. He was also a
senior business counsel charged with intellectual property development and
patent generation. Currently he is on assignment in IBM Systems Group
(previously IBM Server Group). He has given many invited talks and has over
50 publications and 20 patents to his credit.
Download