A 2.5-GHz Low-Power, High Dynamic Range, Self

advertisement
1618
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 40, NO. 8, AUGUST 2005
A 2.5-GHz Low-Power, High Dynamic Range,
Self-Tuned Q-Enhanced LC Filter in SOI
Xin He and William B. Kuhn, Senior Member
Abstract— -enhanced LC filter technology offers a promising
approach to remove the off-chip preselect filter still required in current transceivers. However, reported designs fail to meet stringent
system specifications such as dynamic range and noise figure for
existing wireless standards. The complexity and inaccuracy of frequency and tuning have also prevented acceptance in industry
applications. This paper presents a low-power and high-performance design targeted at Bluetooth in a silicon-on-isolator (SOI)
CMOS process. Drawing 5 mA from a 3-V supply, it achieves 23-dB
voltage gain (14-dB power gain), approximately 6-dB noise figure,
and a 153-dB Hz 1-dB compression point dynamic range, when
operating with a 70-MHz bandwidth at 2.5 GHz. A simplified aptuning is also demonstrated, making
proach to frequency and
-enhanced LC filtering feasible in industry application for the
first time.
Index Terms—Active filters, bandpass filters, full integration, LC
filters, RFCMOS, self-tuning, silicon-on-isolator (SOI), SOS.
I. INTRODUCTION
T
O ACHIEVE lower cost and smaller size, the trend in
wireless communication transceiver design has been
toward full integration from the antenna port to the information
source/sink, eliminating discrete components without compromising performance such as power consumption, selectivity,
noise figure, and dynamic range. However, in current gigahertz-range receivers, a bulky and expensive off-chip bandpass
filter before the LNA is still required to handle the existence of
large out-of-band interference. Great efforts have been made to
use an on-chip -enhanced LC filter to replace such off-chip
preselect filters [1]–[11].
As shown in Fig. 1, a -enhanced filter can be applied to
replace both the discrete preselect filter and the LNA, providing
comparable bandwidth, front-end gain, and net noise figure
performance at similar power consumption [9]. Moreover, it
has been demonstrated that the filter can offer good protection
against out-of-band interferers, since, with proper design in
CMOS, compression occurs first at the filter output and the
filter’s gain decreases outside its passband [5].
While the -enhanced filter approach is not suitable for
frequency-division-duplex (FDD) transceivers without time
Manuscript received August 12, 2004; revised February 8, 2005.
X. He was with the Department of Electrical and Computer Engineering, Kansas State University, Manhattan, KS 66506-5204 USA. He is
now with Philips Research East Asia, Shanghai 200070, China (e-mail:
xin.he@philips.com).
W. B. Kuhn is with the Department of Electrical and Computer Engineering, Kansas State University, Manhattan, KS 66506-5204 USA (e-mail:
wkuhn@ksu.edu).
Digital Object Identifier 10.1109/JSSC.2005.852043
Fig. 1. Receiver front-end designs. (a) Traditional. (b)
Q-enhanced filter.
regulation for transmit and receive, it is suitable for time-division-duplex (TDD) and time-division multiple-access (TDMA)
transceivers as well as receive-only systems. Fortunately, many
existing communication systems support TDD, such as GSM,
Bluetooth, TDD/WCDMA, and WLAN. As data communication dominates future communication systems with the
increasing demand for large volume multimedia transfer, TDD
is becoming more popular due to its high efficiency. If a suitable
on-chip tunable filter is available, it can decrease the size of
future multistandard transceivers significantly by eliminating
different copies of discrete filters for different bands.
Unfortunately, the use of such on-chip bandpass RF filters
has been prevented so far by their limited performance capabilities relative to system specifications. Stringent system requirements such as high compression point, narrow bandwidth,
high dynamic range, and low noise figure have been determined
based on capabilities of off-chip filters [9]. In commercial designs, an LNA typically presents a 2-dB noise figure, 20-dB
gain, and a 1-dB input compression point of about 20 dBm,
while consuming 15 mW [12]. Therefore, the LNA combined
with a 3-dB insertion-loss discrete filter achieves a net 5-dB
noise figure and 17-dBm input compression point with 17-dB
voltage gain, yielding a 1-dB compression point dynamic range
of 152 dB (in a 1-Hz normalized bandwidth). To date, reported
-enhanced filter designs have succeeded in bandlimiting the
signal sufficiently, but fail to match these stringent noise figure
and dynamic range capabilities [8]. The best-documented examples present a noise figure of 19 dB and dynamic range of
137 dB Hz while drawing as much as 50 mW or more [4]–[9],
far from the needed requirements, even for Bluetooth, which has
very relaxed system specifications.
Another obstacle for -enhanced filtering is the need for
tuning. The master–slave (M/S)
automatic frequency and
tuning technique has dominated research endeavors in most
on-chip filter research and applications due to its relative
simplicity compared to self-tuning approaches such as the
correlated tuning loop technique [13] and orthogonal reference
tuning [14]. However, to avoid interference problems, the
master-oscillator frequency must be kept away from the center
0018-9200/$20.00 © 2005 IEEE
HE AND KUHN: 2.5-GHz LOW-POWER HIGH-DYNAMIC RANGE SELF-TUNED
Fig. 2. Simplified
Q-enhanced filter circuit-level diagram.
frequency of the slave filter, which violates the assumption that
master and slave resonators demonstrate the same quality factor
and leads to significant inaccuracy for tuning. This leads to
problems in achieving a predictable filter and bandwidth.
In this paper, we present a low-power, low-noise, high-dynamic-range -enhanced LC filter with a low-complexity
self-tuning technique addressing all of these problems. One
reason for the poor performance in previous designs is that
only low- (less than 10) inductors are available in standard CMOS or BiCMOS processes. The solution is to use an
silicon-on-isolator (SOI) process with high bulk resistivity
such as silicon-on-sapphire (SOS) where high- inductors are
available ( 20 between 2–3 GHz). With the use of on-chip
high-Q inductors and appropriate circuit topology, this filter
achieves 23-dB voltage gain, approximately 6-dB noise figure,
a 1-dB compression point dynamic range of 153 dB Hz, and
70-MHz bandwidth at 2.5 GHz, while drawing only 5 mA
from a 3-V supply. The performance is comparable to that
of a conventional LNA with an off-chip bandpass filter. A
low-complexity self-tuning technique is also demonstrated
which maintains frequency and bandwidth over fabrication
tolerances and temperature.
The paper is organized as follows. Section II presents the
new -enhanced LC filter design at 2.5 GHz. The noise figure
and dynamic range analysis is then given in Section III. In Section IV, we address the self-tuning technique in the RF bandpass
filter. The measured results are demonstrated in Section V. Finally, the paper is wrapped up with the conclusion in Section VI.
II.
-ENHANCED LC FILTER DESIGN
It has been shown that active LC filters can achieve a significantly larger dynamic range than inductorless active filters
- filters with similar power consumption [9], [15].
such as
Therefore, the -enhanced filter is a more promising approach
to meet the stringent performance requirements in wireless receivers. The basic idea is to use negative resistance to boost the
in a lossy LC tank. Fig. 2 shows a simplified second-order
represents equiv(one-pole) bandpass filter topology, where
alent parallel loss resistance from the finite tank at resonance.
realizes the required negative resisNegative conductance
tance to compensate for the loss in the tank. The effective quality
factor of the filter at the resonant frequency can be shown to be
[1]
-ENHANCED LC FILTER IN SOI
1619
. Indeed, the
it can be set as high as desired with appropriate
filter core can be tuned to oscillate if negative transconductance
. This fact, which is
is sufficiently large, i.e., greater than
typically viewed as a problem, can be turned into an asset as
Section IV on tuning shows.
Fig. 3 shows the full schematic of the new -enhanced
filter in SOS targeted at Bluetooth applications. The filter is
composed of an input balun L2, C2, input transconductor
FETs M1a/b, cascode FETs M2a/b, and tank L1, C1, while
the negative resistance is implemented by cross-coupled FETs
M3a/b (plus more switched copies) and analog fine tuning using
source-degenerated negative transconductor M4a/b. Frequency
tuning is similarly provided by switched capacitances and MOS
varactor control.
Two buffers M5a/b are positioned on the differential output of
the filter to isolate the LC tank from the following loads during
test. M5a/b function as source followers with 200- output impedances. A 50- buffer would introduce much bigger parasitic
capacitance and losses, thus degrading the of the LC tank in
the filter, and was not used since 250- probes were available
for test.
A. LC Tank
In a -enhanced LC filter, the base quality factor of the
LC tank before -enhancement is critical for the performance
such as dynamic range and noise figure [15]. Fortunately, high
spiral inductors with
of 20 are becoming available
in RF-targeted SOI processes. RF-compatible SOI processes
such as SOS use a high resistivity or insulating substrate, as
opposed to the very low (0.02 cm) or moderate (20 cm)
resistivity of standard CMOS, therefore reducing or eliminating
substrate losses [16]. As an example, Fig. 4 presents a square
inductor tested in the target process and its measured result. A
six-element frequency-independent model of spiral inductor
in SOI process has been proposed in [16]. This model, which
was used during the design of the filter, includes only losses
from resistance and eddy currents within the metal traces of the
spiral.
In the design of Fig. 3, a 1.4-nH one-turn inductor with a of
approximately 30 is used in the tank. Due to finite- switched
capacitors and MOS varactor, the net of the tank is around 20.
Hence, the target bandwidth of 70 MHz can be achieved with
a moderate enhancement factor of 1.8, allowing good dynamic
range performance to be achieved [9].
B. Input Matching Network
An important innovation in this filter is to adopt an LC input
balun which decreases the noise figure significantly and provides additional preselect filtering [17]. A 6:1 voltage step-up
balun formed by L2 and C2 is employed to convert the unbalanced 50- source to a differential source for differential input
is given by
stage M1a/b. The ideal upconversion ratio
(1)
(2)
where
is the base quality factor of the LC tank, which is dominated by the inductor in the low gigaherz range. Theoretically
dewhere is resonant angular frequency of the balun and
notes the 50- source. Due to the finite for the inductor L2a
1620
Fig. 3. Schematic of
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 40, NO. 8, AUGUST 2005
Q-enhanced filter.
and L2b (around 20 in implementation), each one presents a se; thus the actual ratio is
ries resistance of
(3)
and the transformed differential source impedance seen by
M1a/b is
(4)
This conversion increases the percentage of the total output
noise contributed by the source, therefore decreasing the noise
figure. However, two factors limit the maximum ratio. First,
series resistance increasing with the inductor impedance required in a higher upconversion ratio will degrade noise figure
eventually as it becomes comparable to source impedance.
Second, the size of FETs M1a/b also upper bounds the ratio. In
the input balun, capacitor C2 absorbs the parasitic capacitance
from gate to source and gate to drain in M1a/b to resonate
L2. As the upconversion ratio increases, the capacitance of C2
gets smaller, eventually becoming comparable to the gate input
capacitance. To keep the bandwidth and the noise figure impact
of the input balun reasonable, the fixed capacitance C2 was set
to about five times bigger than the total gate capacitance.
C. Input Transconductors
The differential input transconductors M1a/b use a
grounded-tail configuration to provide maximum possible
signal swing within the amplifier. A current-tail structure
exhibits better common-mode rejection but degrades signal
swing headroom. Since there may be strong out-of-band signals
presented to the filter, bigger signal swing is a more important
concern. M1 uses a special zero-threshold device offered by
the target IC process, both to minimize the noise figure and
to optimize the dynamic range performance in the filter [9],
overdrive at 0.5-V bias
[10]. With the resulting 0.5-V
voltage, signals of up to 1-V peak differential can be tolerated
at the filter input, translating to approximately 9-dBm input
compression at the antenna terminals. Additional protection
comes from the input balun at large frequency offsets. For
example, the calculated out-of-band blocking performance at
500-MHz offset (1.9-GHz PCS band) is 3 dBm, or 7 dB
above the Bluetooth requirement. Transistors M2a/b are also
implemented with zero-volt threshold devices to minimize
noise figure, but are three times larger than M1. This minimizes
the voltage swing at the drains of M1, allowing M1’s gates to
be biased at a lower voltage, maximizing signal swing within
the LC resonator.
D. Negative Transconductance and
Tuning Components
In contrast to conventional -enhanced filters, a current-reuse strategy is employed to implement negative transconductance in this design. In the filter, negative transconductor
FETs M3a/b share the same dc current with input FETs M1a/b,
without sacrificing dynamic range. M3a/b are implemented
with low-noise, regular-threshold (0.8-V) PFET devices to
provide approximately 1-V peak differential signal swing
across the LC tank’s effective resistance at resonance before
leaving the active region (onset of output compression). In the
HE AND KUHN: 2.5-GHz LOW-POWER HIGH-DYNAMIC RANGE SELF-TUNED
Fig. 4.
-ENHANCED LC FILTER IN SOI
1621
A 4.6-nH inductor in the target process. (a) Die photograph. (b) Measurement result.
simulation, around 0.7-V peak differential swing is achieved
at the compression point, translating to a calculated output
compression point of 1.5 dBm in the 350- net LC tank loss
resistance and to an in-band input-referred compression point
of 15.5 dBm at a voltage gain of 23 dB (14-dB power gain).
The PFET negative transconductors are sized to offset losses
within the on-chip spiral inductor. By increasing the negative
transconductance, the filter turns into an oscillator. Due to the
variation of the inductor caused by fabrication and temperature
variation, this negative transconductance has to be tuned to
a proper value to obtain the desired gain and bandwidth for
the filter. Instead of using a fixed-size transconductor with a
single tunable copy, -tuning is implemented by a bank of
switched transconductors, with widths of 80, 40, 20, 10, and
5 m. This bank provides coarse digital control, while analog
is achieved through controlling the
fine-tuning control of
input gate voltage of long-channel FETs M4a/b, which act as
degenerative resistors when operated in triode region.
E. Frequency-Tuning Components
Since the fixed MiM capacitor C1 in the LC tank combined
with parasitic capacitances from connected devices is subject
to fabrication variation of about 10%, some frequency tuning
is required. Usually a MOS varactor is applied to implement
full-range frequency tuning. However, with the very high
inductors used, normal MOS varactors exhibit relatively poor
varactor with
quality factor. For example, combining a
inductor results in a tank of only 12. Another disada
vantage of MOS varactors is variation of quality factor with the
voltage difference across the gate and the drain–source, therefore increasing filter nonlinearity. A better approach is to employ a bank of digital switched capacitors for coarse tuning and a
small MOS varactor for fine tuning. Another benefit of switched
capacitors is that we may use simple digital control algorithm
to achieve the desired frequency, as discussed later.
To illustrate the design considerations in using switched
capacitances, Fig. 5 shows a circuit-level model including
Fig. 5. Conventional switched capacitor and its equivalent on/off model.
parasitic capacitances and on-resistance. The switched capacitor is turned on by setting
low and turned off by setting
high. The turn-on capacitance is the capacitance from the
fixed capacitor, and the corresponding series resistance mainly
which determines
comes from the FET channel resistance
the on-state . When the capacitor is turned off by setting
high, the capacitance is the fixed capacitance in series
and the
with the parasitic gate–drain overlap capacitance
. In this state, the
parasitic layout drain–source capacitance
quality factor is determined by the turn-on resistance of the
bottom nFET in the inverter and the parasitic series resistance
in gate–drain capacitance. Hence, we need to use a large-size
transistor for an inverter to keep high and the full gate–drain
overlap capacitance determines the net off-state capacitance
achieved.
In the implemented design shown previously in Fig. 3,
modified circuitry which increases the average and achieves
lower off-state capacitance is employed. As shown in Fig. 6, a
large-value resistor is added between the inverter and the transistor. The turn-on capacitance is the same as the conventional
digital switched capacitor, but the primary turn-off capacitance
) now becomes the fixed capacitance in series with
(ignoring
and the gate–source
the gate–drain overlap capacitance
. Hence, the off-state capacitance is
overlap capacitance
reduced by one-half, therefore increasing the on–off ratio.
1622
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 40, NO. 8, AUGUST 2005
Fig. 7.
Fig. 6. Proposed switched capacitor and its equivalent on/off model.
Unfortunately the parasitic layout capacitance
between
the source and the drain cannot be neglected. To obtain the parasitic drain–source capacitance in the switch FET , we first
measured the off capacitance of from the gate to the drain and
the source, with the gate biased at zero voltage and drain–source
. Therefore,
and
connected together, denoted by
were given by
(5)
Next. we measured the off capacitance between the drain and
source with the gate floated, denoted by . Since
consists of
the two series capacitance
and
together with the parallel
was calculated by
Simplified single-ended equivalent circuit for noise analysis.
, body-effect conductance since it is absent is SOI processes,
and flicker noise which has no effect for small-signal inputs
within the filter passband. First we investigate the output
noise contributed from different noise sources in the filter. The
small-signal model of half the differential filter with thermal
noise is presented in Fig. 7. From the small signal model, the
, series resistance
noise sources include 50- source
from the inductors in the input balun, parallel resistance
of half LC tank, and channel thermal noise current
in the input transconductor, channel thermal noise current
in the cross-coupled output transconductor. The output
noise generated by the current noise in the cascode transistor
can be neglected. The reason is that the noise current sees
from the source assuming an infinite
an impedance of
drain-to-source resistance, thus inducing an opposite current in
the channel which cancels out the noise current.
The following equations present the single-sided output noise
, and
, respectively:
contributed by
(6)
(9)
The measurement shows that
is comparable to half of the
in the target process. Nevertheless, the new apoff-state
proach still achieves higher Q and moderately improved on/off
ratio compared to conventional one.
(10)
(11)
(12)
III. NOISE FIGURE AND DYNAMIC RANGE ANALYSIS
(13)
For a -enhanced filter to replace a discrete preselect RF filter
and subsequent LNA as shown in Fig. 1, the noise figure must be
minimized. In this section, we derive an approximate expression
for the noise figure which leads to the tradeoffs involved to yield
the desired noise performance and dynamic range.
The voltage gain of the filter is composed of two parts: the
in the input balun and the voltage gain
upconversion ratio
in the filter core.
is described as
(7)
where
is the half-circuit input transconductance and
denotes the effective parallel resistance in the half LC tank enhanced by negative transconductance.
can be estimated as
In (9) and (10), the noise is divided by 2 as it is split between
the two half-circuits in the balun. In (12) and (13), an additional
comes from injection of noise from the opposite
factor of
half-circuit (since the noise is uncorrelated with this half-cirand
are the noise factors associated with
and
cuit).
, respectively, which are generally
in short-channel tranof the intrinsic
sistors. In the target process, the measured
NMOS used for input transconductance is about 2, and the meaof regular PMOS used for negative transconductance
sured
exhibits the lowest value with around 1 [17].
can be written as
Now the total rms output noise
(14)
(8)
where
denotes the effective quality factor in the LC tank,
is the half LC tank’s parallel loss resistance, and
is the
base quality factor of the LC tank.
is defined as the
-enhancement factor.
A complex thermal noise model for CMOS transistors is
adopted in [18]. In our derivation, for simplicity, we neglect
since it can be absorbed by
drain-to-source conductance
and the noise factor is expressed as
(15)
Substituting (9)–(14) into (15) yields (16), shown at the bottom
of the next page. The calculated result matches the simulation
of HP ADS within less than 5% inaccuracy by using
HE AND KUHN: 2.5-GHz LOW-POWER HIGH-DYNAMIC RANGE SELF-TUNED
, which is the noise factor in the long-channel transistor.
When the input balun operates as an ideal balun, i.e.,
and
, (16) can be simplified to (17), shown at the bottom
of the page. To achieve lower noise factor and better noise rejection performance, several tradeoffs are involved in the filter
and power consumpdesign. The first one exists between
will
tion. From (16), increasing the input transconductance
decrease the noise factor of the filter. However, to handle large
interference signals, we have to keep the dc overdrive voltage
conon transconductor M1a/b at about 0.5 V, hence higher
sumes more power. The second tradeoff is between series resisand voltage upconversion ratio in the input balun as
tance
discussed in previous section. The optimum ratio can now be
found from (16).
Once the noise figure is known, dynamic range (DR) can
be calculated based on supported voltage swings at the filter
output. While many definitions for dynamic range exist, a
simple but useful one is given by the ratio of 1-dB input compression point to the input-referred noise floor relative to 1-Hz
bandwidth, given by
(18)
corresponding to
With respect to output peak voltage
1-dB output compression point in this filter, it can be written as
(19)
Substituting (16) into (19) yields (20), shown at the bottom of
the page, where denotes the enhancement factor
.
A simplified DR defined by the ratio of 1-dB output compression point to the output noise floor in a second-order -enhanced filter has been shown in [15]
(21)
where
represents the output noise floor,
denotes the
final IF bandwidth, and is the noise factor of transistors used
-ENHANCED LC FILTER IN SOI
1623
is proportional to the enhancein the filter. Note that
ment factor from (7) and (8), so that (20) also presents the
appearing in (21).
penalty of
To compare circuits independent of bandwidth, the 1-dB
compression DR relative to a 1-Hz bandwidth is applied to
define a figure of merit (FOM) [9]
(22)
where
is the dc power. To date, the best FOM achieved
is 121 dB Hz/mW. The filter reported here achieves
141 dB Hz/mW through the use of high
inductors and
careful attention to noise figure issues. This value is comparable to a traditional front-ends built with a discrete preselect
filter followed by a comparable-power LNA [9].
IV. SELF-TUNING TECHNIQUE
As previously discussed, traditional M/S tuning is difficult to implement in -enhanced filters and consumes extra
power and chip area. Self-tuning offers a better alternative for
TDD/TDMA systems. However, traditional self-tuning suffers
from the complexity of an additional reference generation
circuit and the feedthrough caused by the switches between the
signal path and reference signal. In this paper, we introduce a
low-complexity self-tuning which eliminates the need for RF
reference signal and switches, as shown in Fig. 8. The essential
in the filter until
concept of this approach is to increase
it oscillates and then recover a filter with desired frequency
and bandwidth from oscillation by decreasing appropriately
negative transconductance inside the circuitry. The full tuning
procedure is proposed as follows.
1) Turn the filter into an oscillator by increasing the control signal.
2) Tune the oscillator frequency to the desired center frequency of the filter.
3) Decrease the -tuning control and detect the transition
point when oscillating dies away.
(16)
(17)
(20)
1624
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 40, NO. 8, AUGUST 2005
Fig. 9.
Fig. 8. Low-complexity self-tuning technique eliminating RF reference and
switches.
4) Back off the control by an additional fixed amount to
establish the desired final bandwidth.
In this filter, the frequency tuning can be realized by a simple
digital tuning algorithm with the help of a classic phase-frequency detector (PFD) which functions as a frequency detector
here. The output of the PFD is applied to a counter controlling
the 4-b switched-capacitor bank within the filter. First we initialize the count to the lowest frequency and maximum amplitude by turning on all 4-b switched capacitors and 5-b negative transconductance. During tuning, if the frequency after divide-by- is lower than the reference signal to the PFD, we
decrease one LSB in each reference clock. The tuning is continued until the signal frequency after divide-by- is higher
than the reference signal, indicated by the inverted PFD output.
Assuming a moderate 1-MHz clock at the counter, this will
require only a maximum of 16 s which is negligible in the
system. Subsequently, the -tuning is performed to set the filter
bandwidth. Note that the shift between the oscillating frequency
and the center frequency of filter is small because the tank cirvalue is high and the parasitic capacitance variation due
cuit
to turning on/off of cross-coupled FETs is small compared to
fixed capacitance in the LC tank. This is an additional benefit of
using high inductors within the design. Moreover, we avoid
the use of a full PLL circuit here since the center frequency does
not need to be highly accurate.
For
tuning, the following issues must be addressed.
• A transition point must be accurately detected when oscillation dies away by decreasing negative transconductance.
• A constant or predictable negative transconductance variation between the filter with desired bandwidth and the
transition point of oscillator must be possible in the tuning
frequency range and commercial temperature range.
• The minimum tuning time should be far shorter than the
idle slot between transmit and receive in the target communication system.
Fig. 9 plots the curve of the oscillation amplitude versus width
of cross-coupled FETs with 0.5- m length in the oscillator. The
FET’s size where the oscillating amplitude reaches zero is defined as the ideal transition point in this paper. From (1), the
Oscillating amplitude versus negative transconductor width.
theoretical negative transconductance at the transition point can
be expressed as
(23)
From (1), the required cross-coupled transconductance
is
(24)
and thus the variation of negative transconductance between the
oscillator and the filter is found from
(25)
Since the inductance of
is relatively immune to fabrication
and temperature variation in a good RF process with high ,
is fixed for a given
.
We may consider the transistor size corresponding to a small
amplitude, for example, one fifth of the maximum amplitude as
the approximate transition point, which can be easily detected
by a comparator working with an RF amplitude detector. The
small difference between idea transition point and approximate
.
transition point can be factored into
In actual implementation,
is realized by turning off a
pair of cross-coupled transistors with a fixed size. Unfortunately,
the negative transconductance provided by the fixed pair may
vary with temperature. The transconductance of a long channel
transistor is expressed as
(26)
The work [19] indicates that the temperature coefficients for
and
are about 5000 ppm/ C and mV C
,
on
are opposite in
respectively. Since the effects of and
of
can be reduced to
sign, the temperature coefficient
1500 ppm/ C by choosing overdrive voltage
between
is
0.6 and 0.3 V. From (1), the temperature coefficient of
given by
(27)
HE AND KUHN: 2.5-GHz LOW-POWER HIGH-DYNAMIC RANGE SELF-TUNED
Fig. 10.
-ENHANCED LC FILTER IN SOI
1625
Simplified oscillator model.
As the enhancement factor is limited to 2, 70 C temperature
variation around 10%, which is acceptchange results in
able in real applications. Furthermore, some first-order temperature compensation circuits may be possible by careful design
variation with temperature, thus allevito further mitigate
variation.
ating
Finally, we investigate the time issue in this tuning technique.
The total tuning time consists of the time consumed by the setup
of the oscillator, frequency tuning, and transition point detection, in addition to the settling time after turning the oscillator
back into the desired filter.
A simplified oscillator model is depicted in Fig. 10. The
transfer function for this model is expressed as
(28)
With a zero input, the envelope of oscillation corresponding to
the elapsed time is given by
(29)
reaches the stable amplitude. The oscillating setup
before
when suftime is quite short due to the small time constant
ficient excess negative transconductance is used. In simulation,
the setup time is less than 100 ns, negligible compared to system
idle time in existing communication systems. The analysis for
time cost in transition point detection is quite complicated. We
decrease the negative inductance by turning off a constant-size
transconductor at each step, thus increasing the negative resistance and the time constant in (29). However, we can estimate
the time constant as ten times the initial time constant for the
worst case, and the detection time can be on the order of microseconds by ignoring the digital switching time in searching
the transition point.
Frequency-tuning time in the oscillator is mainly determined
the by the frequency of the reference signal in the PFD. Due to
the removal of feedback, the response time to achieve the desired frequency is significantly faster than that of a traditional
PLL. For quiescent time from the oscillator to the filter, the response time is also short due to the small time constant
in the desired filter. For example, with
and
pf in the half LC tank after -enhancement, the time constant
is about 2 ns.
From the analysis above, the accumulated tuning time can
be limited on the order of microseconds, which is significantly
smaller than the idle time in existing TDD and TDMA communication systems.
Fig. 11.
Die photograph.
V. MEASURED RESULTS
The filter described above has been implemented in a 0.5- m
SOS process with one poly layer and three metal layers. A photograph of the chip is shown in Fig. 11. The area consumed by
the filter is approximately 2.5 mm . The one-turn spiral inductor
positioned on the right of the die is the inductor in the LC tank.
The single spiral inductor on the left of the die is the tail inductor. Two parallel spiral inductors next to the tail inductor are
used for the input balun, in which the layout polarity of inductors is chosen to cancel coupling to other inductors.
The design targeted a tuning range from 2.25 to 2.65 GHz to
span the Bluetooth band and provide for tuning. The measured
frequency range is from 2.45 to 2.85 GHz, due to the difference in inductor values between EM simulator and fabrication.
In measurements of noise figure and dynamic range, the frequency controls were set to 2.55 GHz and the desired selectivity
was achieved through manual setting of the coarse controls.
Automatic tuning was then implemented with the aid of bench
test equipment for measuring frequency and amplitude and software for adjusting filter tuning controls.
Measured current consumption is 5 mA (excluding buffer)
at a 3-V power supply. The actual filter response (S21) versus
-tuning and frequency tuning are shown in Figs. 12 and 13,
respectively.1
The filter response over the full range of 0–6 GHz is shown in
Fig. 14. Ultimate rejections are 25 dB at most frequencies of
interest (cellular/PCS/cordless). At low frequency, the observed
rejection is believed be limited by the single-sided measurement
at the output and/or insufficient supply bypassing.
The hot/cold source method with a 15-dB ENR source was
applied to measure the noise figure of the filter. The differential
voltage gain of the filter was set to 23 dB (14-dB power gain)
by coarse control, obtaining a bandwidth of 70 MHz
. A 4.9-dB noise floor variation at resonance was observed,
which gives a noise figure of 11.8 dB for the total cascaded
stages including the attenuating probe and spectrum analyzer
measurement system, translating to a noise figure of 5.8 dB for
1These measurements were obtained with an HP8753E network analyzer
using direct probing with a model ECP18 Picoprobe on the filter input and a
model-10 250-
(5:1 voltage attenuating) passive probe on the buffer circuit
output. Raw measurements showed a 5-dB gain after S21 calibration with both
probes connected to a “through” substrate. Actual filter voltage insertion gains
shown in the plots include an adjustment of +18 dB due to 6-dB loss from
the output buffer voltage division, 6 dB from the single-sided sampling of the
output, and approximately 6 dB due to the open-circuit source value presented
to the high-impedance port-2 probe during calibration.
1626
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 40, NO. 8, AUGUST 2005
Q-tuning.
Fig. 12.
Actual filter response versus
Fig. 13.
Actual filter response versus frequency tuning.
Fig. 15. Noise floor spectrums in hot/cold source conditions at 1% fractional
bandwidth (2.541-GHz center frequency, 150-MHz span; 40-dBm reference,
and 10 dB per division).
0
Fig. 16.
Improvement of the compression point outside the filter bandwidth.
TABLE I
SUMMARY OF MEASURED PERFORMANCE
Fig. 14.
Filter response over 0–6 GHz.
the filter.2 Operating at high gain and high mode, the filter
presents slightly a higher noise figure, but the measurement can
be more accurate because the gain appearing after the probe
exhibits a few decibels instead of loss in the low gain mode. As
2An error analysis indicates that the actual noise figure may be up to 1 dB
higher or lower than the measured value due to tolerances in the measurement.
shown in Fig. 15, a noise floor power variation of 7.3 dB was
recorded with a voltage gain of 32 dB and 27-MHz bandwidth,
translating to a noise figure of 7 dB when operating as a tunedpreselect filter with a one percent fractional bandwidth
.
The measured input 1-dB compression point is 15 dBm at
a voltage gain of 23 dB at 70-MHz bandwidth, resulting in a
normalized DR of 153 dB Hz and a FOM of 141 dB Hz/mW. If
HE AND KUHN: 2.5-GHz LOW-POWER HIGH-DYNAMIC RANGE SELF-TUNED
-ENHANCED LC FILTER IN SOI
1627
TABLE II
PERFORMANCE COMPARISONS BETWEEN LNA AND INTEGRATED RF FILTERS IN THE LITERATURE
Fig. 17.
Q-tuning over 80-MHz frequency range.
a spectrum analyzer and Labview software. From the GPIB interface, the computer reads the frequency and amplitude of the
oscillator and changes the setting of the negative transconductance to find the transition point. A simple constant variation of
FET size between the transition point and the filter with of
about 40 was found to be viable in an 80-MHz frequency range,
as shown in Fig. 17. Fig. 18 demonstrates the actual filter response achieved in over a temperature range of 50 using the
proposed technique.
A summary of measured performance is provided in Table I
for the cases of operation as a fixed-tuned prselect/LNA and a
tuned-RF preselect/LNA system block. Table II presents comparisons between this design operating in the fixed-tuned mode
and other reported designs. Note that, in the closest published
result [7], there is no gain and the actual FOM would be significantly degraded (at least 10 dB lower) if the LNA stage is
included.
VI. CONCLUSION
Fig. 18.
Q-tuning over 50
temperature variation at 2.55 GHz.
used at the narrower 27-MHz bandwidth, the compression point
is 23 dBm due to the higher associated gain.
As shown in Fig. 16, the compression point improves outside
the filter bandwidth as the filter gain falls. This plot was produced at the nominal 70-MHz filter bandwidth. Once the compression point reaches about 6 dBm, the improvement slows
as the input transconductor begins to enter the nonlinear region
exceeding
overdrive).
of operation (
To validate the basic functionality of the self-tuning technique, automatic tuning has been performed with the help of
The design described in this paper implements a practical
one-pole -enhanced filter in a 0.5- m SOI process. With the
help of high inductors and an on-chip voltage step-up input
balun, it achieves a FOM significantly exceeding that of previous designs. The measured FOM is 12 dB higher than the best
reported result, mainly due to the relatively low noise figure
in this design. The filter provides 23-dB voltage gain (14-dB
power gain), a noise figure of 6 dB, and a normalized 1-dB compression point DR of 153 dB Hz when operating at a 70-MHz
bandwidth centered at 2.5 GHz, with 5-mA current and 3-V
supply. The result is comparable to that of existing commercial front-end designs employing an off-chip bandpass filter and
LNA. A simple but efficient self-tuning methodology is also introduced in the paper, which eliminates the need for complex
circuits and interacting frequency and tuning loops in conventional master–slave or self-tuning techniques. With the demonstrated performance and simple tuning circuits, -enhanced filtering is shown to be practical for industry applications.
ACKNOWLEDGMENT
The authors would like to thank D. Nobbe and D. Kelly
at Peregrine Semiconductor for collaboration on process and
1628
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 40, NO. 8, AUGUST 2005
inductor issues and A. Orsborn for background work on
noise issues.
REFERENCES
[1] W. B. Kuhn, F. W. Stephenson, and A. Elshabini-Riad, “A 200-MHz
CMOS Q-enhanced LC bandpass filter,” IEEE J. Solid-State Circuits,
vol. 31, no. 8, pp. 1112–1122, Aug. 1996.
[2] S. Pipilos, Y. P. Tsividis, J. Fenk, and Y. Papananos, “A Si 1.8-GHz
RLC filter with tunable center frequency and quality factor,” IEEE J.
Solid-State Circuits, vol. 31, no. 10, pp. 1517–1525, Oct. 1996.
[3] R. Duncan, K. W. Martin, and A. S. Sedra, “A Q-enhanced active-RLC
bandpass filter,” IEEE Trans. Circuits Syst. II, Analog Digit. Signal
Process., vol. 44, no. 5, pp. 341–347, May 1997.
[4] W. Gao and W. M. Snelgrove, “A linear integrated LC bandpass filter
with Q enhancement,” IEEE Trans. Circuits Syst. II, Analog Digit. Signal
Process., vol. 45, no. 5, pp. 635–639, May 1998.
[5] W. B. Kuhn, N. K. Yanduru, and A. S. Wyszynski, “Q-enhanced LC
bandpass filters for integrated wireless applications,” IEEE Trans. Microw. Theory Tech., vol. 46, no. 12, pp. 2577–2586, Dec. 1998.
[6] D. Li and Y. Tsividis, “Design techniques for automatically tuned integrated gigahertz-range active LC filters,” IEEE J. Solid-State Circuits,
vol. 37, no. 8, pp. 967–977, Aug. 2002.
[7] T. Soorapanth and S. S. Wong, “A 0-dB IL 2140 30 MHz bandpass
filter utilizing Q-enhanced spiral inductors in standard CMOS,” IEEE J.
Solid-State Circuits, vol. 37, no. 5, pp. 579–586, May 2002.
[8] F. Dulger, E. Sanchez-Sinencio, and J. Silva-Martinez, “A 1.3-V 5-mW
fully integrated tunable bandpass filter at 2.1 GHz in 0.35-m CMOS,”
IEEE J. Solid-State Circuits, vol. 38, no. 6, pp. 918–928, Jun. 2003.
[9] W. B. Kuhn, D. Nobbe, D. Kelly, and A. W. Orsborn, “Dynamic range
performance of on-chip RF bandpass filters,” IEEE Trans. Microw.
Theory Tech., vol. 37, no. 10, pp. 685–694, Oct. 2003.
[10] W. B. Kuhn, “Fully integrated bandpass filters for wireless transceivers—Problems and promises,” in Proc. IEEE Midwest Symp.
Circuits and Systems, 2002, pp. 69–72.
[11] X. He and W. B. Kuhn, “A fully integrated Q-enhanced LC filter with
6 dB noise figure at 2.5 GHz in SOI,” in Proc. IEEE Radio Frequency
Integrated Circuits (RFIC) Symp., 2004, pp. 643–646.
[12] MBC13916 General Purpose SiGe:C RF Cascode Amplifier. Motorola,
Data Sheet.
[13] T. L. Brooks and P. M. VanPeteghem, “Simultaneous tuning and signal
processing in integrated continuous-time filters: The correlated tuning
loop,” in Proc. IEEE Int. Symp. Circuits and Systems, 1989, pp. 651–654.
[14] W. B. Kuhn, A. Elshabini-Riad, and F. W. Stephenson, “A new tuning
technique for implementing very high Q, continuous-time, bandpass filters in radio receiver applications,” in Proc. IEEE Int. Symp. Circuits and
Systems, 1994, pp. 5-257–5-260.
[15] W. B. Kuhn, F. W. Stephenson, and A. Elshabini-Riad, “Dynamic range
of high-Q OTA-C and enhanced-Q LC RF bandpass filters,” in Proc.
IEEE Midwest Symp. Circuits and Systems, 1994, pp. 767–771.
[16] W. B. Kuhn, X. He, and M. Mojarradi, “Modeling spiral inductors in
silicon-on-sapphire processes,” IEEE Trans. Electron Devices, vol. 51,
no. 5, pp. 677–683, May 2004.
6
[17] A. W. Orsborn, “Noise analysis and automatic tuning of Q-enhanced LC
bandpass filters,” M.S. thesis, Dept. Electr. Comput. Eng., Kansas State
Univ., Manhattan, KS, 2001.
[18] BSIM3v3 Manual, 1999. Univ. of Calif., Berkeley.
[19] Y. Wu, C. Shi, M. Ismail, and H. Olsson, “Temperature compensation
design for a 2.4 GHz CMOS low noise amplifier,” in Proc. IEEE Int.
Symp. Circuits and Systems, vol. 1, Geneva, Switzerland, May 2000, pp.
323–326.
Xin He received the B.S. and M.E. degrees in
instrumentation from the University of Science and
Technology of China, Hefei, China, in 1997 and
2000, respectively, and the M.S. and Ph.D. degrees in
electrical engineering from Kansas State University,
Manhattan, in 2002 and 2004, respectively.
In the summer of 2004, he joined Philips Research
East Asia, Shanghai, China, as a Senior Research
Scientist. His research interests include RF filter
design, high-efficiency power amplifiers, PLLs, and
modeling of spiral inductors and varactors in SOI
processes.
William B. Kuhn (S’78–M’79–SM’98) received the
B.S. degree in electrical engineering from Virginia
Polytechnic Institute and State University (Virigina
Tech), Blacksburg, in 1979, the M.S. degree in electrical engineering in 1982 from the Georgia Institute
of Technology (Georgia Tech), Atlanta, and the Ph.D.
degree from Virginia Tech in 1996.
From 1979 to 1981, he was with Ford Aerospace
and Communications Corporation, Palo Alto,
CA, where he designed radio receiver equipment
including frequency synthesizers and bit synchronizers. From 1983 to 1992, he was with the Georgia Tech Research Institute,
Atlanta, working primarily in radar signal analysis and mixed-signal circuit
simulator development. In 1996, he joined Kansas State University, Manhattan,
as an Assistant Professor, later becoming an Associate Professor in 2000. He
currently teaches courses in communications theory, radio and microwave circuit/system design, and VLSI. His research is primarily targeted at low-power
radio electronics in CMOS, BiCMOS, and SOI technologies.
Dr. Kuhn was the recipient of the Bradley Fellowship in 1993 from Virginia
Tech and a Faculty Early Career Development (CAREER) Award from the National Science Foundation in 1999. At Kansas State he has also been the recipient of the Hollis Award for Excellence in Undergraduate Teaching in 2001, the
Eta Kappa Nu Distinguished Faculty Award in 2002 and 2003, and the Paslay
Professorship in Electrical and Computer Engineering in 2004.
Download