River 1
River 1
Introduction to Wireless
Communication Circuits
Introduction to Wireless
Communication Circuits
Introduction to Wireless
2nd Edition
2nd Edition
Series Editors
Indexing: All books published in this series are submitted to Thomson Reuters Book
Citation Index (BkCI), CrossRef and to Google Scholar.
Forouhar Farzaneh
Professor
Sharif University of Technology, Iran
Ali Fotowat
Associate Professor
Sharif University of Technology, Iran
Mahmoud Kamarei
Professor
University of Tehran, Iran
Ali Nikoofard
Research Engineer
University of California at San Diego, USA
Mohammad Elmi
Research Engineer
KavoshCom Asia Co., Iran
River Publishers
Published, sold and distributed by:
River Publishers
Alsbjergvej 10
9260 Gistrup
Denmark
River Publishers
Lange Geer 44
2611 PW Delft
The Netherlands
Tel.: +45369953197
www.riverpublishers.com
c 2020 River Publishers
As wireless technology takes over every aspect of our lives, the university
curricula must keep up with the developments and impart proper skills to their
graduates so as to prepare them for this rapidly-evolving industry. In particular,
the vast body of undergraduate students must be trained in this domain, but
efficiently, as course proliferation is undesirable in most universities.
“Introduction to Wireless Communication Circuits” addresses this need by
selecting the most relevant topics and teaching them in a language that appeals
to undergraduate students. The textbook methodically guides the reader through
the concepts and, using numerous detailed examples, reenforces these concepts.
The reader is then invited to exercise his/her understanding by solving problems
at the end of each chapter.
The contents of the book have been chosen carefully to allow coverage in
one semester or quarter. That is, the book can serve as a self-contained text
that the students can read “cover to cover” in one term without skipping any
major sections. These pedagogical aspects of the book facilitate its use for both
students and instructors.
Behzad Razavi
Professor
University of California, Los Angeles
February 2018
v
Preface to the Second Edition
During the past couple of years where we used this book as our teaching
reference in wireless communication circuits, we encountered a number of
points to be clarified or improved in the text. To this effect, we have prepared
the material for the second edition. Scores of equations and a few figures were
added; therefore, the second edition contains 1161 equations and 505 figures
which help more in the analysis and understanding of the text. We have changed
the text in hundreds of instances to make the material straightforward and more
comprehensible. We hope that this new edition will be more useful for the
students and practicing engineers.
vii
Preface to the First Edition
ix
x Preface to the First Edition
Foreword. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
. . the
Preface to . . . Second
. . . . . . .Edition
. . . . . . . . . . . . . . . . . . . . . . . . . . . vii
. . the
Preface to . . . First
. . . .Edition
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xix
List of Figures
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxxvii
List of Tables
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxxix
List of Abbreviations
I Part 1
1 The Amazing World of Wireless Systems . . . . . . . . . 3
1.1 Introduction to Communication Circuits . . . . . . . . . . . . . 3
1.2 Signal Levels and Rayleigh Fading . . . . . . . . . . . . . . . . 8
1.3 Calculation of the Sensitivity in Different Standards . . . . . 9
1.4 Considerations in RF System Design . . . . . . . . . . . . . . . 10
1.5 A Basic Understanding of Frequency Synthesizers . . . . . 17
1.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.7 References and Further Reading . . . . . . . . . . . . . . . . . 21
1.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
xi
xii Contents
2 Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.1 An Introduction to Oscillators . . . . . . . . . . . . . . . . . . . 27
2.2 First Approach: Positive Feedback . . . . . . . . . . . . . . . 27
2.3 Second Approach: Negative Resistance/
Conductance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.4 Oscillator Topologies . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.4.1 Common-Emitter Oscillator Circuit . . . . . . . . . . . . . . . 36
2.4.2 Common-Base Oscillator Circuit . . . . . . . . . . . . . . . . . 37
2.4.3 Common-Collector Oscillator Circuit . . . . . . . . . . . . . . 37
2.4.4 Colpitts versus Hartley Oscillators, a New Insight . . . . . . 38
2.5 Crystal Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.5.1 Datasheet of a Family of Crystals . . . . . . . . . . . . . . . . 45
2.6 Calculation of the Oscillation Frequency Including
the Device Parasitics . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.7 Quality Factor of Reactive Elements . . . . . . . . . . . . . . 47
2.8 Nonlinear Behavior in Amplifiers . . . . . . . . . . . . . . . . . 49
2.9 A Note on the Modified Bessel Functions of the
First Kind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
2.10 Large-Signal Transconductance and Harmonic Tuned
Amplifiers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.10.1 Case I: Resonant circuit is tuned to the first harmonic
of the input frequency (tuned amplifier case) . . . . . . . 58
2.10.2 Case II: Resonant circuit is tuned to the second
harmonic of the input frequency (frequency multiplier
case) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.11 Differential Bipolar Stage Large-Signal
Transconductance . . . . . . . . . . . . . . . . . . . . . . . . . . 63
2.12 Inductive and Capacitive Dividers (Impedance
Transformers) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
2.12.1 Tapped Capacitive/Inductive Impedance
Transformers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
2.13 Analysis of Large-signal Loop Gain of an Oscillator . . . . 74
2.13.1 Increasing the Quality Factor and the Frequency
Stability with a Crystal . . . . . . . . . . . . . . . . . . . . . . . . 77
2.13.2 Oscillator Harmonics Calculation . . . . . . . . . . . . . . . . 81
2.14 Colpitts Oscillator with Emitter Degeneration . . . . . . . . 82
2.15 MOS Stage Large-Signal Transconductance . . . . . . . . . 83
2.16 Differential MOS Stage Large-Signal
Transconductance . . . . . . . . . . . . . . . . . . . . . . . . . . 86
2.17 An Oscillator With a Hypothetical Model . . . . . . . . . . . 89
2.18 A MOS Oscillator with Differential Gain Stage . . . . . . . . 90
Contents xiii
II Part 2
3 PLL, FM Modulation, and FM Demodulation . . . . 127
3.1 Frequency Modulation . . . . . . . . . . . . . . . . . . . . . . . 127
3.2 Frequency Demodulation . . . . . . . . . . . . . . . . . . . . . 130
3.2.1 Phase Detector . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
3.2.2 Gilbert Cell as a Phase Detector . . . . . . . . . . . . . . . . 131
3.2.3 Quadrature Phase (FM) Detector . . . . . . . . . . . . . . . 134
3.3 Basics of PLLs and their Application as an FM
Demodulator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
3.3.1 The Transfer Function of the First-Order PLL . . . . . . . . . 144
3.4 Further PLL Applications . . . . . . . . . . . . . . . . . . . . . . 152
3.4.1 FM with PLL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
3.4.2 PLL Application in Frequency Synthesizers and
Its Transfer function . . . . . . . . . . . . . . . . . . . . . . . . . 158
3.5 Advanced Topic: PLL Type II . . . . . . . . . . . . . . . . . . . 161
3.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
3.7 References and Further Reading . . . . . . . . . . . . . . . . 165
3.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
4 Mixers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
4.1 Mixer Concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
4.1.1 The Conceptual Behavior of Single-Diode Mixers . . . . 169
4.1.2 A Nonlinear Circuit as a Mixer . . . . . . . . . . . . . . . . . . 170
4.2 Third Order Intermodulation Concept in a Nonlinear
Amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
4.2.1 Characteristic of Third-Order IM and Measurement
Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
4.3 Basic Concept of Third-Order IM in a Basic Mixer . . . . 174
4.3.1 The Desired Channel Blocking with the Third-Order
IM Component . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
xiv Contents
III Part 3
7 Transmission Lines and Impedance Matching . . 283
7.1 An Introduction to Radio-Frequency Amplifiers
in Receivers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
7.1.1 Transmission Line . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
7.2 Wave propagation Equations in Transmission Line
for R = 0 and G = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . 286
7.2.1 General Wave Propagation Relations in lossy
Transmission Lines . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
xvi Contents
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 469
xix
xx List of Figures
Figure 6.22 (a) Passive (spiral) inductor model. (b) The active
MOS inductor topology and its equivalent circuit . . . . . 275
Figure 6.23 Decreasing device input capacitance via series feedback . 277
Figure 6.24 The possible feedback leakage path and the oscillation
issue in multistage amplifiers . . . . . . . . . . . . . . . . 278
Figure 7.1 A generic RF front-end . . . . . . . . . . . . . . . . . . . 284
Figure 7.2 The model of a high-frequency amplifier . . . . . . . . . . 284
Figure 7.3 The phase and gain response of a lossless T-line . . . . . . 285
Figure 7.4 Impedance transformation property of the T-line, (a) The line
terminated by 50 Ω exhibits an impedance of 50 Ω through-
out the line, (b) The line terminated by an unmatched load
can exhibit both an inductive or capacitive impedance seen
through it depending on the position, (c) Transformation of
a short circuit to an open circuit using a quarter wavelength
T-line, and (d) Load impedance inversion using a quarter
wavelength T-line . . . . . . . . . . . . . . . . . . . . . . 285
Figure 7.5 The lumped model of a differential transmission line for a
differential length ∆Z . . . . . . . . . . . . . . . . . . . . 286
Figure 7.6 A transmission line divided into consecutive differential
lumped sections . . . . . . . . . . . . . . . . . . . . . . . 289
Figure 7.7 Characteristic impedance of a line . . . . . . . . . . . . . 292
Figure 7.8 A terminated transmission line . . . . . . . . . . . . . . . 294
Figure 7.9 The terminated T-line . . . . . . . . . . . . . . . . . . . . 296
Figure 7.10 A terminated T-line to the intrinsic impedance . . . . . . . 298
Figure 7.11 A short-terminated T-line . . . . . . . . . . . . . . . . . . 299
Figure 7.12 The input impedance of a short-terminated T-line . . . . . 299
Figure 7.13 The input impedance of an open-terminated T-line . . . . 300
Figure 7.14 The input impedance of an open-terminated T-line . . . . 301
Figure 7.15 A T-line with its load and source impedances . . . . . . . 301
Figure 7.16 Equivalent model of Figure 7.15 for input impedance
calculation . . . . . . . . . . . . . . . . . . . . . . . . . 302
Figure 7.17 (a) Demonstration of a T-line with its load and source
impedances, and (b) transmission of the wave back and forth
on the line . . . . . . . . . . . . . . . . . . . . . . . . . . 303
Figure 7.18 Implementation of a narrowband inductor and a narrowband
capacitor using a short and an open-terminated line . . . . 305
Figure 7.19 (a) Implementation of a parallel inductor and a parallel capac-
itor using short-circuit and open-circuit stubs, respectively,
(b) Implementation of a series inductor using a short-length
high-impedance transmission line, (c) Implementation of a
parallel capacitor using a short-length low-impedance stub,
all in microstrip technology . . . . . . . . . . . . . . . . . 306
Figure 7.20 A loaded resonant circuit . . . . . . . . . . . . . . . . . . 307
Figure 7.21 The effect of source impedance on Q . . . . . . . . . . . . 308
Figure 7.22 The effect of inductance on Q . . . . . . . . . . . . . . . 308
Figure 7.23 Output power characteristic for a DC circuit . . . . . . . . 308
Figure 7.24 The general case of maximum power transmission to the load 309
xxx List of Figures
Figure 9.3 Output stability circles on the source plane Smith chart for
five cases (the shaded areas correspond to the stable region),
(a) for |S22 | > 1 and conditional stability, (b) for |S22 | < 1,
the stability circle intersecting the chart while not comprising
the chart center, and consequently, conditional stability, (c)
for |S22 | < 1, the stability circle intersecting the chart while
comprising the chart center, and consequently, conditional
stability, (d) for |S22 | < 1, the stability circle does not in-
tersect the chart and consequently, unconditional stability,
and (e) for |S22 | < 1, the stability circle comprises the whole
chart, and consequently, unconditional stability . . . . . . 383
Figure 9.4 Input and output reflection coefficients on the Smith chart . 388
Figure 9.5 Steps for input matching on the Smith chart . . . . . . . . 389
Figure 9.6 Steps for output matching on the Smith chart . . . . . . . 390
Figure 9.7 The overall matching network . . . . . . . . . . . . . . . 390
Figure 9.8 The load matching using a series capacitor and a parallel
inductor to achieve 9 dB gain . . . . . . . . . . . . . . . . 393
Figure 9.9 Source matching . . . . . . . . . . . . . . . . . . . . . . 394
Figure 9.10 The input and output matching networks for complex source
and load impedances . . . . . . . . . . . . . . . . . . . . 395
Figure 9.11 (a) The input stability circle and the corresponding source
reflection coefficient. (b) The output stability circle and the
chosen load reflection coefficient . . . . . . . . . . . . . . 397
Figure 9.12 The noise model of a two-port amplifier . . . . . . . . . . 399
Figure 9.13 Model for noise figure calculation in a cascade of
two stages . . . . . . . . . . . . . . . . . . . . . . . . . . 401
Figure 9.14 Constant NF contours on the plane of input reflection
coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . 405
Figure 9.15 Constant NF and the normalized available power gain
contours on the input reflection coefficient plane at 6 GHz 406
Figure 9.16 Input matching on the Smith chart . . . . . . . . . . . . . 407
Figure 9.17 Output matching on the Smith chart . . . . . . . . . . . . 408
Figure 9.18 The overall matching network . . . . . . . . . . . . . . . 409
Figure 9.19 Constant power gain and constant NF contours for the given
transistor on the Smith chart . . . . . . . . . . . . . . . . 411
Figure 9.20 Equivalent circuit model of the transistor in Example 9.8 . 412
Figure 9.21 The π matching network employed in Example 9.8 . . . . 412
Figure 9.22 The amplifier circuit for determining various power gains . 416
Figure 9.23 Cascaded amplifiers to determine the overall S-parameters 416
Figure 9.24 A transistor cascaded by either of series or parallel
resistances . . . . . . . . . . . . . . . . . . . . . . . . . . 417
Figure 9.25 A transistor amplifier with corresponding load and source
impedances and the bias circuitry . . . . . . . . . . . . . 417
Figure 9.26 The Smith chart to design an amplifier with specific NF
and GA . . . . . . . . . . . . . . . . . . . . . . . . . . . 418
Figure 9.27 The cascade amplifiers/mixer for determination of the overall
noise figure . . . . . . . . . . . . . . . . . . . . . . . . . 419
xxxiv List of Figures
Figure 9.28 A cascade of amplifiers for the output power calculations . 419
Figure 9.29 The transistor MOS amplifier for input and output matching 419
Figure 9.30 The two-stage amplifier to be designed using a BGU7007
internally matched LNA . . . . . . . . . . . . . . . . . . 420
Figure 9.31 The constant noise and the constant available gain circles at
the source plane of the transistor at 4 GHz . . . . . . . . . 421
Figure 9.32 The equivalent circuit of a FET transistor and the associated
matching circuits . . . . . . . . . . . . . . . . . . . . . . 422
Figure 9.33 A cascode MOS stage amplifier . . . . . . . . . . . . . . 422
Figure 9.34 The low-noise amplifier with the corresponding input and
output matching circuits . . . . . . . . . . . . . . . . . . 423
Figure 9.35 The measured source admittances for a noise figure F = 3
(source admittance plane) . . . . . . . . . . . . . . . . . . 423
Figure 10.1 A graphical representation of the compromise between the
linearity and the efficiency in a typical power amplifier . . 426
Figure 10.2 Typical compression curve of a power amplifier . . . . . . 427
Figure 10.3 Compression curve of the PAE of an amplifier . . . . . . . 428
Figure 10.4 Power amplifier followed by a matching network . . . . . 428
Figure 10.5 PDF of the transmitted power of the PA for a mobile set
in an urban or suburban area . . . . . . . . . . . . . . . . 429
Figure 10.6 The effect of the TX-band noise at the receiver in a
full-duplex system . . . . . . . . . . . . . . . . . . . . . 430
Figure 10.7 A PA with the PA driver and the interstage matching
networks . . . . . . . . . . . . . . . . . . . . . . . . . . 431
Figure 10.8 Typical compression of the output power versus the input
power of a PA . . . . . . . . . . . . . . . . . . . . . . . . 431
Figure 10.9 Typical gain compression in a PA . . . . . . . . . . . . . 431
Figure 10.10 (a) P1dB and IP3 points in a nonlinear system, (b) Output
spectrum in a nonlinear system with a two-tone input . . . 432
Figure 10.11 An amplifier with a nonlinear capacitance at its load . . . 433
Figure 10.12 A typical AM to PM characteristics for a nonlinear
amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . 434
Figure 10.13 Output spectrum of Figure 10.11 due to PM conversion,
(a) k 1, and (b) k < 1 or k ≈ 1 . . . . . . . . . . . . . . 434
Figure 10.14 Generation of undesired harmonic and intermodulation com-
ponents at the output of a nonlinear PA . . . . . . . . . . . 435
Figure 10.15 A nonlinear tuned amplifier corresponding to its nonlinear
transconductance . . . . . . . . . . . . . . . . . . . . . . 436
Figure 10.16 Spectral regrowth in a nonlinear PA as seen in the frequency
domain . . . . . . . . . . . . . . . . . . . . . . . . . . . 437
Figure 10.17 Spectral regrowth in a nonlinear PA, single tone shown in
the time domain . . . . . . . . . . . . . . . . . . . . . . . 437
Figure 10.18 ACPR definition in CDMA (IS-95) standard . . . . . . . . 438
Figure 10.19 Degradation of in-band SNR due to nonlinearity . . . . . . 439
Figure 10.20 Representation of error vector magnitude (EVM) in a
sampling I − Q constellation . . . . . . . . . . . . . . . . 440
List of Figures xxxv
Figure 10.21 A comparison of the input and the output constellation for
a 16-QAM modulation for a nonlinear power amplifier in
between . . . . . . . . . . . . . . . . . . . . . . . . . . . 440
Figure 10.22 Class A power amplifier . . . . . . . . . . . . . . . . . . 441
Figure 10.23 (a) Quiescent point determination in class A power amplifiers
and (b) Drain voltage, drain current, and device power loss
of the circuit depicted in Figure 10.22 . . . . . . . . . . . 442
Figure 10.24 A class B power amplifier with the corresponding output
current waveform . . . . . . . . . . . . . . . . . . . . . . 443
Figure 10.25 (a) The quiescent operating point and the corresponding
waveforms for the class B PA and (b) The output current, the
output voltage, and the power loss waveforms for a single-
device class B power amplifier . . . . . . . . . . . . . . . 444
Figure 10.26 A class B push–pull power amplifier with the output
current shown as a summation of the positive and the negative
half-cycle currents . . . . . . . . . . . . . . . . . . . . . 445
Figure 10.27 Class AB PA’s quiescent point and the corresponding current
waveform . . . . . . . . . . . . . . . . . . . . . . . . . . 446
Figure 10.28 A class C power amplifier and its corresponding output
current waveform . . . . . . . . . . . . . . . . . . . . . . 447
Figure 10.29 The quiescent point of a class C amplifier and the
corresponding output current, output voltage, and the
power loss waveforms . . . . . . . . . . . . . . . . . . . 447
Figure 10.30 The output current waveforms alongside the efficiencies
and the power capabilities corresponding to different
classes of power amplifiers as a function of the conduction
angle, α . . . . . . . . . . . . . . . . . . . . . . . . . . . 449
Figure 10.31 A typical class D power amplifier topology and the
corresponding voltage and current waveforms . . . . . . . 449
Figure 10.32 A typical circuit topology and current waveforms of a
class 1/D power amplifier . . . . . . . . . . . . . . . . . . 451
Figure 10.33 A typical class E power amplifier topology and the
corresponding voltage and current waveforms . . . . . . . 452
Figure 10.34 A typical class F power amplifier topology and the
corresponding voltage and current waveforms . . . . . . . 453
Figure 10.35 A typical class F power amplifier topology using a
quarter-wavelength transmission line . . . . . . . . . . . . 453
Figure 10.36 Block diagram of a class S power amplifier and its
corresponding waveforms . . . . . . . . . . . . . . . . . 454
Figure 10.37 A typical class sinusoidal to PWM signal converter . . . . 454
Figure 10.38 Different linearization methods in power amplifiers . . . . 456
Figure 10.39 A typical output power versus input power curve in decibels
depicting the input back-off and the output back-off with
respect to the saturation point . . . . . . . . . . . . . . . . 456
Figure 10.40 A schematic view of predistortion technique . . . . . . . 457
Figure 10.41 The block diagram of an adaptive predistorter technique . 457
xxxvi List of Figures
xxxvii
List of Abbreviations
xxxix
xl List of Abbreviations
2 Oscillators . . . . . . . . . . . . . . . . . . . . . . . 27
2.1 An Introduction to Oscillators
2.2 First Approach: Positive Feedback
2.3 Second Approach: Negative Resistance/Conductance
2.4 Oscillator Topologies
2.5 Crystal Oscillators
2.6 Calculation of the Oscillation Frequency Including
the Device Parasitics
2.7 Quality Factor of Reactive Elements
2.8 Nonlinear Behavior in Amplifiers
2.9 A Note on the Modified Bessel Functions of the First Kind
2.10 Large-Signal Transconductance and Harmonic Tuned
Amplifiers
2.11 Differential Bipolar Stage Large-Signal Transconductance
2.12 Inductive and Capacitive Dividers (Impedance Trans-
formers)
2.13 Analysis of Large-signal Loop Gain of an Oscillator
2.14 Colpitts Oscillator with Emitter Degeneration
2.15 MOS Stage Large-Signal Transconductance
2.16 Differential MOS Stage Large-Signal Transconductance
2.17 An Oscillator With a Hypothetical Model
2.18 A MOS Oscillator with Differential Gain Stage
2.19 Voltage-Controlled Oscillators
2.20 Special Topic: Nonlinear Device Fed by Sinusoidal Large-
Signal
2.21 Datasheet of a Voltage-Controlled Oscillator
2.22 Conclusion
2.23 References and Further Reading
2.24 Problems
ʌ/2
÷N VCO
÷M
Wireless system and circuit design is one of the most interesting fields in electrical
engineering. From the economic point of view, wireless applications can be categorized
into cellular/smart phones, cordless phones, wireless data networks, sensor networks,
global positioning systems, and digital television broadcasting (terrestrial or satellite
based). A huge investment has already been made in this sector and experts project
further growth in the years to come. From the engineering point of view, the design
of wireless systems has different levels of abstraction which are relevant to radio
frequency (RF) antennas, wave propagation phenomena, RF and microwave circuit
design, evaluation of noise and intermodulation phenomena, digital modulation, coding,
and digital signal processing.
Receiver 5 7
I
1 2 3 4
Q
9 10 11 12 13
π/2 6 8
÷N VCO
15
PLL ÷M
14
16 18 Transmitter
I 26
21 22 23 24 25
20
+
Q
PLL
17 π/2 19
Figure 1.1: The general block diagram of a transceiver (radio transmitter plus
receiver).
aid radios, and WiFi networks, we need to select the frequency channel of interest
properly (i.e., reception) while interferences from all the other systems may be present.
Similarly, we must transmit a channel in the frequency allocated to an application
without causing excessive interference for other applications (i.e., transmission).
To start after the receiving antenna, we normally use a band-pass filter, block (1),
to preselect the spectrum of our application (e.g., the full 25 MHz bandwidth in the
869.2–893.8 MHz receive band of GSM). In block (2), the weakly received signal
will be amplified, usually by about 5 to 20 dB. This block normally consumes several
milli-amps of current because it is normally operating in class A and at the highest
frequency. In block (3), the amplified signal is downconverted through a mixer which
brings the signal to a lower frequency for further processing. Block (3) is symbolized
by a multiplication sign because the multiplication of two sinusoids is known to result
1.1 Introduction to Communication Circuits 5
in two new frequency components at the sum and the difference frequencies. Block
(4) is a band-pass filter that usually eliminates the undesired frequencies and selects
one of the two output signals. The frequency at block (4) is called the intermediate
frequency (IF). The channel selection in radios is performed by changing the local
oscillator (LO) frequency applied to block (3). The frequency from block (4) onward
is fixed, making its processing more simple. Blocks (5) and (6) downconvert the IF to
the baseband (in modern radios, where blocks (3) and (4) are suppressed and the RF
frequency is directly converted to the baseband, they are called zero-IF downconverter).
Blocks (7) and (8) are low-pass filters that are narrow enough to select the desired
information. The I and Q outputs go to a DSP (digital signal processing) block for
further digital processing intended for the display or the speaker for example. The
difference between filters (1), (4), (7), and (8) is that as we go through the receiving
chain, the filters become narrower, eliminating undesired frequency components. To
generate the LO signal used to derive block (3), we start with a crystal oscillator in
(9) whose frequency is usually between 5 MHz and 50 MHz. Block (10) divides the
crystal oscillator frequency by integer N to provide a lower stable frequency. Block
(13) is a voltage-controlled oscillator (VCO) whose output frequency is divided by
integer M in block (14). The resulting two frequencies out of block (10) and (14) are
compared in block (11). The output of block (11) is low-pass filtered by block (12)
which provides an error voltage to drive the VCO. The ratio M is digitally controlled.
When the loop is settled, the frequency of the VCO will be set to M/N of the frequency
of the crystal oscillator. We describe these blocks in more detail in the upcoming
chapters. It is important to note at this point that the first LO generates the signal
required by block (3) to select the desired channel. Block (15) is a fixed oscillator that
supplies the second LO for blocks (5) and (6).
For the transmitter portion in modern receivers, the baseband signals (either voice,
video, or data), after analog-to-digital conversion, form the I and Q signals which are
low-pass filtered by blocks (16) and (17). The outputs are upconverted by blocks (18)
and (19) mixers and summed in block (20). The resulting signal is band-pass filtered
in block (21) which is called the IF of the transmitter, then applied to a second mixer
of block (22), and is upconverted to the desired RF channel. Filter (23) is a band-pass
filter that selects the desired radio frequency and leaves out the undesired components.
Block (24) is a power amplifier that may amplify the output to the desired wattage.
Block (25) is the final stage filtering that will guarantee proper compliance with the
regulatory standard preventing undesired frequency components (here, the harmonics
or the intermodulation) for other systems or subscribers. The LO frequencies needed
for the transmit path might be generated by the same scheme as the receiver. The
difference between filters (16), (17), (21), (23), and (25) is that as we move forward in
the transmit path, they become wider to allow the transmission of the full spectrum of
the application to be used. For example, in GSM 850, the final filter (25) is a band-pass
filter in the range of 824.2 MHz–849.2 MHz.
As an example for a transceiver, we investigate the block diagram of the second
generation (2G) Digital AMPS system (DAMPS1 ). We have deliberately chosen this
system because it includes both analog and digital modulations. In this system, the
channel spacing is 30 kHz. The AMPS standard was fully analog, but evolved to
1 Digital advanced mobile phone system.
6 Chapter 1. The Amazing World of Wireless Systems
contain digital modulation in DAMPS. The analog modulation in this system is based
on frequency modulation with a maximum frequency deviation of 12 kHz (and given the
3 kHz baseband, consequently a total bandwidth of 30 KHz). The digital modulation
is based on π/4 QPSK. Considering a possible bit rate of about 60 kb/s in each
channel with a bandwidth of 30 KHz, three users can be present. The increased number
of subscribers is due to digital modulation and the time division among them, and
consequently sequential transmission of digital information. The procedure of sharing
one channel between three users is based on three time slots (time division multiple
access or TDMA). Each subscriber’s speech data are recorded and transmitted in
its time slot. This is accomplished at the cost of a maximum of three time slots
delay. The speech data are also compressed with advanced algorithms to reduce
its bit rate (to less than 20 kb/s). It is possible that the subscribers are in different
geographical positions, and as a result, we need a base station for management and
control of the three time slots allocated to different subscribers in different places. In
this system, the frequencies of the receive and the transmit have 45 MHz difference.
The frequency allocation for this system is in the range of 824 MHz−849 MHz which
is used for transmission of the mobile set and is called uplink. Similarly, the downlink
for this system (the reception frequency of the mobile set) is defined in the range of
869 MHz−894 MHz that is used by the base station. As it is evident, the difference
between the center frequencies of the downlink and the uplink bands is 45 MHz and
each has a 25 MHz bandwidth. In Figure 1.2, the spectrum usage and the frequency
allocation of this system are shown.
In Figure 1.2, two simultaneous subscribers are shown. In Figure 1.3, it is shown
that there is a free 30 kHz channel between two adjacent channels.
As illustrated in Figure 1.3 even though the bandwidth of each channel is 30 kHz,
in the same cell, 60 kHz channel spacing is considered. This is due to maintaining
an interference-free reception that is discussed in the following chapters. What was
described earlier for DAMPS can be similarly applied for GSM2 assuming a 200 kHz
bandwidth and digital performance. It is recommended that in mobile networks there
be always an empty channel between two adjacent channels.
6SHFWUXP
7UDQVPLW 5HFHLYH
EDQG EDQG
I0+]
6SHFWUXP
N+]
N+]
5HFHLYH
EDQG
I0+]
1HLJKERULQJ
8VHU
$GMDFHQW
Figure 1.3: The DAMPS channel spacing stipulates one empty channel between
two adjacent channels in every cell (30 kHz guard band is considered between
two adjacent channels).
where Z0 is the reference impedance. It is observed that once both signals have the
2 2
7 3 2 7 3
1 7 3 1
6 4 1 6 4
5 6 4 5
5
Figure 1.4: Three clusters of seven cell frequency distribution for DAMPS or
GSM.
8 Chapter 1. The Amazing World of Wireless Systems
same reference impedance, the power ratio and the voltage ratio in dB (decibels) would
have the same value. Another popularly used definition in radio engineering is dBm
which is used for describing the absolute power of the signals and is defined as the
ratio of the power in milli-watts to a 1 mW reference power and defined as
PmW
P( dBm) = 10 log (1.2)
1 mW
We now explain the difference between dB and dBm. When we use the term dB, we are
expressing the logarithmic ratio of two signal amplitudes; once we are using dBm, we
are expressing the logarithmic power ratio of the signal with respect to a 1 milli-watt
reference or describing the power in dBm. Here are a few conversion examples in
Equation 1.3.
5 dBm = 3 mW (1.3a)
0 dBm = 1 mW (1.3b)
− 10 dBm = 0.1 mW (1.3c)
− 100 dBm = 0.1 pW (1.3d)
With the above definitions, the sensitivity in the GSM system implying the minimum
signal which could be properly detected is about −103 dBm, and for DAMPS, the
sensitivity is −114 dBm.
35G%
5HODWLYH
UHFHLYHG
VLJQDO
VWUHQJWK
'
'LVWDQFHIURP
WUDQVPLWWLQJDQWHQQD
field strength. This model predicts the amplitude of the signal which experiences
changes or fading by passing through the multipath media. The distribution of this
statistical model is based on Rayleigh distribution. The received signal strength in
general is dependent on the base station’s transmitted power, its antenna gain, and
the fading phenomenon in the propagation medium. In general, the received signal
strength diminishes with distance plus or minus some local variations. In large cities,
the above-mentioned problem becomes more acute because of the reflections from the
ground, and reflections or diffractions from the multiple buildings on the propagation
path. Due to higher traffic in large cities and higher population density, the number of
cells is increased which in turn results in lower cells’ radii. In suburban or rural areas,
where there are not too many base stations, the radius of the cell may increase and the
RF power coming from the base station may be as high as 43 dBm and the minimum
detectable signal may be as low as −103 dBm. The difference between the transmitted
and the received signal being as high as 146 dB implies the complexity of design at
radio frequencies. Handling the large dynamic range required in the rural or suburban
areas and the significant Rayleigh fading phenomenon encountered in urban areas is
one of the main challenges encountered in radio systems.
signal-to-noise ratio. Now rewriting Equation 1.4 in decibels, one (at the standard
temperature of 290◦ k) obtains
As such, the sensitivity in these three standards will become as what is demonstrated
in Table 1.1.
* G% $XGLR
%: N+] DPSOLILHU
IF N+]
9RLFH
VLJQDO
* G% * G%
)UHTXHQF\GHPRGXODWRU
Figure 1.6: System-level schematic of the DAMPS receiver (the analog portion).
a few advantages as follows. First of all, achieving high gain is much easier at low
frequencies rather than high frequencies. Secondly, in most receivers the IF frequency
is held constant allowing for more accurate narrowband filter/amplifier design. Up
to the mixer, the signal has experienced 4 dB loss in the passband of the filters and
12 dB gain of the RF amplifier (8 dB total gain) which for an input of 0.5 µVrms
brings the signal to 1.25 µVrms level. In a mixer circuit, the analog multiplication
occurs for which we assume an ideal multiplier. Therefore, the sum and difference
components of 882 + 972 = 1854 MHz and 972 − 882 = 90 MHz are generated. The
former component is eliminated in the first band-pass filter, while the latter reaches the
next stage. At this point, the desired signal is further amplified by 7 dB in the active
mixer and then injected into the filter with 60 kHz bandwidth.
The desired channel passes alongside adjacent channels that are now attenuated.
Finally, the second mixer with an LO frequency of 90.455 MHz brings the 90 MHz
signal to 455 kHz with, say, 12 dB gain. The 455 kHz signal goes through two 30 kHz
filters plus a chain of amplifiers with 40 dB and 60 dB gain. It is instructive to note that
most of the gain is obtained at low frequency and with a small current consumption. In
addition, the design of narrowband band-pass filters is generally much easier at lower
frequencies rather than high frequencies.
A receiver which exploits one downconversion is traditionally called heterodyne
receiver and if more than one downconversion occurs, it is called superheterodyne
receiver. The words heterodyne and superheterodyne, while having historical signifi-
cance, imply one mixing stage, and two or more mixing stages, respectively, and the
prefix hetero- stands for mixing of different frequencies and the word dyne stands
for analog multiplication or mixing. In the full receiver chain of Figure 1.6, the over-
all gain is of the order of 121 dB (the point in the superheterodyne receiver is that
the total amplification is performed in three different frequency ranges, and there-
12 Chapter 1. The Amazing World of Wireless Systems
fore the probability of instability is reduced). If the effective input signal is of the
order of 0.5 µVrms, the overall chain gain of 121 dB brings the low signal level up
to 562 mVrms. The detector shown is a frequency demodulator which detects the
frequency deviation and extracts the voice signal which is applied to the speaker after
audio amplification. The values shown in this example are typical values in the receiver.
Although the lowest gain was placed at the front-end in the low-noise amplifier, it
draws the highest current from the supply voltage (typically near 3 mA). However, the
high gain of 100 dB in the second IF can be achieved with only 750 µA bias current.
This point indicates one of the challenges of high-frequency amplifier design.
It should be noted, however, that the last audio amplifier stage draws a high current
of several tens of milliamperes for the power amplification of the audio signal as
well.
Example 1.1 Is it possible to add more low-noise amplifier stages at the front-
end instead of filtering in order to have a better noise performance?
Answer:
The answer is no. In fact, placing more low-noise amplifiers at the front-end
results in a better noise performance if there were no strong out-of-band interferers.
However, the strong blocker signals without filtering will bring the last low-noise
amplifier stages into saturation which results in decreased effective gain and pos-
sibly the mixing of the desired channel signal with the amplified blocker ones
(this is discussed in more detail in Chapter 4). In addition, the power consumption
cost of high-frequency amplification and the possibility of parasitic feedback may
jeopardize the stability of the front-end (in case of high gain at a single frequency).
The first band-pass filters attenuate out-of-band blocker signals and the first image
signal (2 f0 − fs ), while fs is the RF signal frequency and f0 is the first LO frequency.
As such, the frequency of the first image is 1062 MHz, i.e., 90 MHz above the first LO
frequency. Thus, the importance of those front-end band-pass filters is now obvious.
The third filter after the mixer passes the desired channel, but will attenuate the adjacent
and neighboring channels to some extent. Figure 1.7 shows a possible condition of
the received signals. The two adjacent channels shown in Figure 1.7 will produce
another signal which is due to the mixer nonlinearity and is called the third-order
intermodulation product (IM3 ) which is thoroughly discussed in Chapter 4. The IM3
component will fall on the desired signal; if the IM3 signal is larger than the desired
signal, signal detection will not be possible. Therefore, the third filter mitigates this
issue by attenuating the adjacent and neighboring channels.
Now, we discuss the spectral behavior of the receiver described earlier. As stated
earlier, the importance of the third filter in the receiver is the attenuation of the two
adjacent channels that lie in 60 kHz and 120 kHz away from the desired channel.
This issue is demonstrated in Figure 1.7 in which the weak desired signal lies at the
882 MHz frequency.
Now, consider the frequency response of the third band-pass filter which is shown
in Figure 1.8. Our goal is to calculate the attenuation of the adjacent and the neighbor-
ing channels’ signals if the filter has a second-order behavior.
One may remember that with two poles in the transfer function, the magnitude
of the signal will decrease by 40 dB/decade or 12 dB/octave, or in other words, the
magnitude response will fall by 12 dB once the relative frequency is doubled. Actually,
the attenuation of the filter is calculated based on the relative offset from the center
frequency. For instance, in Figure 1.8, the adjacent channel is just one octave above the
3 dB cut-off frequency; thus, it experiences 12 dB attenuation. Similarly, the neighbor-
ing channel is 120 kHz offset from the desired channel frequency; thus, it experiences
N+] N+]
6LJQDO
SRZHU
G%P G%P G%P
I0+]
Figure 1.7: The desired channel, the adjacent channel, and the neighboring
channel in a typical DAMPS radio signal.
N+] N+]
6LJQDO
SRZHU
G%P G%P G%P
N+]
I0+]
Figure 1.8: Typical received DAMPS signal levels after the third band-pass
filter at 90 MHz .
14 Chapter 1. The Amazing World of Wireless Systems
24 dB attenuation. It can be shown that the attenuation for an nth order filter can be
calculated as
2( f − f0 )
L = 20n log
(1.6)
BW
In Equation 1.6, f0 is the center frequency of the filter, and BW = fU,3 dB − fL,3 dB is
the frequency that the magnitude response of the filter will experience 3 dB attenuation.
In fact, this filter acts as a first channel selection filter. It also attenuates the adjacent
and the neighboring channels; nonetheless, those unwanted channels could be still
stronger than our desired channel. The fourth band-pass filter bandwidth is precisely
equal to one channel bandwidth. Thus, by the second downconversion, both channel
selection and amplification are realized at the second IF. The effect of the fourth filter
is illustrated in Figure 1.9.
Now, consider the attenuation of the fourth filter. The frequency content residing
at 515 kHz and its counterpart residing at 575 kHz will experience 24 dB and 36 dB
attenuation, respectively. Thus, the unwanted adjacent and neighboring channels are
further attenuated. Then, the linear power amplification at low frequency can be
performed by a small current (e.g., 300 µA for 40 dB gain). The fifth filter has the
same behavior as the fourth one. The signal after the fifth filter is demonstrated in
Figure 1.10.
Finally, by the fifth filter, the adjacent channel will be attenuated by 24 dB and
the neighboring channel experiences by another 36 dB attenuation. The spectra of the
wanted and unwanted signals at the FM detector input are depicted in Figure 1.11.
As it is obvious from Figure 1.11, the power of the adjacent channel is below the
desired channel and the neighboring channel has been practically suppressed. Note
that the effect of the additional gains of the second IF amplifiers has been included
in the computation of the final signal levels and the amplifiers are considered to be
linear. In Figure 1.11, we have assumed the total IF gain (90 MHz and 455 kHz)
as G2 = −2 + 12 − 2 + 40 − 2 + 60 = 106 dB which is the sum of gains starting
from the 90 MHz IF all the way to the end of the receiver’s second IF. If we add
G1 = −2 + 12 − 2 + 7 = 15 dB which is the gain of the RF front-end, a total gain of
121 dB for the desired channel is achieved.
N+] N+]
6LJQDO
SRZHU
G%P G%P G%P
N+]
I0+]
Figure 1.9: Typical received DAMPS signal levels after the fourth band-pass
filter at 455 kHz.
1.4 Considerations in RF System Design 15
N+] N+]
6LJQDO
SRZHU
G%P G%P G%P
N+]
I0+]
Figure 1.10: Typical received DAMPS signal levels after the fifth band-pass
filter at 455 kHz.
N+] N+]
6LJQDO
SRZHU
G%P G%P G%P
I0+]
Figure 1.11: Typical received DAMPS signal levels at the input of the frequency
demodulator.
M
fVCO = fcrystal (1.10)
N
The circuit in Figure 1.12 is a negative feedback loop. If the input of the phase detector
is a signal with 30 kHz fundamental, the other input must have the same frequency
component at the steady state. By allowing enough loop gain, the error signal will
tend to zero. This system is called a phase-locked loop frequency synthesizer. We will
go over the concept of the PLL in Chapter 3. It is noteworthy that in a PLL at locked
state, not only will the two input frequencies be the same, but also the phases of the
two signals will track each other with a constant phase offset. Using a programmable
counter, one may change the oscillation frequency to receive the desired channels. For
instance, with a VCO frequency of 972 MHz, we calculate the division ratio as
&U\VWDO
RVFLOODWRU 3KDVH /RZSDVV
GHWHFWRU ILOWHU
9RXW
·1 9&2
'LYLGH
0+] E\1
·0
'LYLGH
E\0
Figure 1.12: Typical integer N frequency synthesizer for the DAMPS receiver.
18 Chapter 1. The Amazing World of Wireless Systems
'$036WUDQVPLWWHU
3RZHUFRQGLWLRQLQJ 5; 7;,7;,
,4
7; 7;47;4
02' 3RZHU$03
9DULDEOH $WWHQ 9 9
DQG JDLQ VZLWFK
6<17+
$QDORJGLJLWDO
FDOO 6ZLWFK
UHJXODWLRQVZLWFK 3$&9
)UHTXHQF\V\QWKHVL]HU
'XSOH[HU /1$
FRQWURO 7%'
0+]
)002' 92XW
)UHTXHQF\V\QWKHVL]HU IRUWUDQVPLWWHU
7;/23// 0+]
%$77(5<
93/& &RQWUROOHU
&21752/ DQG
75$&.+2/'
';
0+] 35(6(7 '&WR'& 9UHJXODWHG
3$B21 FRQYHUWRU 9UHJXODWHG
5 ࢥ'
/1$
3$&9 0+]
1 ';VZLWFK
0+]
6<17+&/.
(17;3//
6<17+'$7
0+]
î08/7
IRUUHFHLYHU
N+]
5 ࢥ'
5;5;,
,4
GHPRG
5;45;4
1
1$ &(5$0,&
'(02'/2 ),/7(5
$*&/9/
0RGXOXVFRQWURO
566,
$8',2'$7$
6<17+&/.
0+]
6<17+'$7
(18+)
0+]
WR3//V
&(5$0,&
),/7(5
$)& N+]
0+]WR
EDVHEDQG
0+]9&7&;2 4XDG
FRLO
'$036UHFHLYHU
14.4 × 106
N= = 480 (1.11)
30 × 103
972 × 106
M= = 32400 (1.12)
30 × 103
To tune to a desired frequency, usually the divider N is constant and the desired channel
is obtained by changing the value of M.
Let’s now proceed to investigate a more complete DAMPS radio as shown in
Figure 1.13.
As depicted in the red portion of Figure 1.13, the signal is received via an antenna.
It is amplified by a low-noise amplifier and the whole band goes through the band-pass
filter. It also attenuates the first image signal. Next, the signal is downconverted by a
mixer to the first IF. Channel selection filter attenuates side-band channels, and by a
second mixer, the signal is translated to a second IF. The LO frequency for a second
mixer comes from a multiplier circuit which makes multiple 6 of crystal frequency.
In the second IF, the signal again is filtered and via two paths goes for digital and
analog demodulation. In the upper path, the signal is demodulated digitally through
an AGC and an I/Q demodulator, and the other path demodulates signal by an FM
1.5 A Basic Understanding of Frequency Synthesizers 19
Example 1.3 A radio receiver has the block diagram shown in Figure 1.14 and
its specifications are denoted on the figure.
625kHz
G=-2dB G=-2dB I
BW=25MHz BW=25MHz BW=6.25MHz
fc=912.5MHz fc=912.5MHz fc=182.5MHz
900-925MHz 625kHz
1.25MHz step G=12dB Q
Wideband 720-740MHz
1MHz step
π/2
(a) The input filters’ specifications are the same. If the desired signal has a
power level of −100 dBm and an image signal accompanies it with a power of
−45 dBm (both at the input of the first RF filter), calculate the out of band atten-
uation of the RF filters, for the image frequency, such that the image signal goes
10 dB below the desired signal at the first mixer input.
(b) Calculate M for channel spacing of 1.25 MHz in the desired band. Note that the
second LO frequency is not fixed.
(c) If at the receiving channel of 912.5 MHz there exists an adjacent channel signal
at 915 MHz, determine the order of the low-pass filters such that the adjacent
channel is rejected by 40 dB. Assume that f3 dB for the low-pass filter is 625 kHz.
Solution:
(a) The image frequency can be obtained as follows
As the RF frequency can be varied between 900 and 925 MHz, then the lower edge
of the receive band is downconverted by 720 MHz. The image frequency would be
Now, consider the upper edge of the receive band. In this case, 925 MHz signal is
downconverted by 740 MHz LO. Thus, the image frequency can be written as
As we desire that the image frequency to be 10 dB lower than the desired signal,
it must be attenuated by 65 dB (55 + 10 = 65 dB). As a result for the identical RF
filters, each one must have an out-of-band attenuation of 32.5 dB at least. The
normalized frequency difference of the image signal will be
log 28.2
D= = 4.81 octaves (1.17)
log 2
32.5
n= = 1.12 (1.18)
6×D
Therefore, we choose n = 2 for the RF filters. This will satisfy the required
attenuation for the other image frequency (540 MHz ) as well.
(b) The channel spacing at the RF frequency in this receiver is 1.25 MHz ; however,
in this architecture 1 MHz spacing is realized by the first mixer (because the crystal
frequency is divided by 15) and 250 kHz is realized by the second mixer (because
the VCO frequency is divided by 4). Thus, the frequency synthesizer will have
a 1 MHz frequency step. As a result, the minimum value of M is 720 and its
maximum value is 740.
(c) As the adjacent channel is 915 − 912.5 = 2.5 MHz above the desired channel,
the attenuation for the nth order low-pass filter can be written as
2.5 MHz
20n log = 40 (1.19)
625 KHz
Then, n = 3.32. So we choose n = 4 as an integer and the filter will be a 4th order
one.
1.6 Conclusion 21
1.6 Conclusion
The world of wireless communications has conquered many aspects of the modern
human life. The technical aspects of this field are of great importance for an electri-
cal/electronic engineer. In this chapter, we made a general presentation for the RF
communication systems. We surveyed the general architecture of an RF transmitter
and an RF receiver. Specifically, we briefly studied the architecture of a superhetero-
dyne receiver which consists of two frequency conversion (mixer) stages as well as a
zero-IF receiver which consists of a receiver with the same LO and RF frequencies.
Furthermore, we observed how in a superheterodyne receiver the large interfering
signals in the adjacent and the neighboring channels are suppressed (or attenuated)
with respect to the desired signal along the receiver chain. In addition, we saw how
using a phase-locked loop and frequency dividers we can synthesize the desired local
frequencies in a receiver. A DAMPS transceiver block diagram was studied as an
example. We deliberately ignored some more complex issues such as the nonlinearity
of the mixers, VCOs, or amplifier circuits here. But as the reader goes forward through
the text, he/she would gain more understanding about the nonlinearity issues. Then the
reader is urged to return back to this chapter to gain more understanding of the related
problems.
1.8 Problems
Problem 1.1 Figure 1.15 shows a triple downconversion receiver in which the input
signal range at the antenna is from 0.02 GHz to 5 GHz. If the first VCO with its initial
control voltage is oscillating at 7.5 GHz,
1. Find the values of M and N in such a way that desired frequencies are provided
for the corresponding mixers.
2. What is the frequency of the image signal at the first mixer’s input? How this
component is eliminated in this structure? What is the frequency of the second
image at the second mixer’s input, and what is the corresponding values at the
input of the antenna. Moreover, find the third image frequency at the third
mixer’s input and its corresponding frequencies alongside the structure.
3. Suppose that the input frequency is 3.5 GHz, then find the first VCO frequency.
In this situation, if the input low-pass filter is a second-order one with 3 dB
frequency of 5 GHz, and if the three subsequent band-pass filters’ frequency
response are as those depicted in Figure 1.16, and if a strong blocker signal
emerges at 100 MHz above the input signal, how much it will be attenuated
through the receiver chain?
22 Chapter 1. The Amazing World of Wireless Systems
VCO VCO ×4
20.0MHz
÷N ÷M
0 5GHz
f 2.5GHz
f 100MHz
f 20.0
f
MHz
100MHz 5MHz
2MHz
Problem 1.2 In the transceiver depicted in Figure 1.17, first IF frequency resides at
90.1 MHz and the second IF is at 455 kHz. If fVCO1 = 966.3 MHz, find the receiving
channel frequency. In this situation, find the frequency of the second VCO such that
the transmitted carrier signal is 45 MHz lower than the received signal.
Amp.
90.555MHz
BW=25MHz
fc=836MHz
Input data
Mod.
VCO1
VCO2
966.3MHz
Problem 1.3 In the FM transceiver depicted in Figure 1.18, determine the unknown
VCO frequencies alongside with division ratio M2 . If fVCO2 and M2 have two different
possible values each, discuss the advantages and disadvantages of either of the values
if the transceiver has a tuning bandwidth of 15 MHz.
IF 0+] IF 0+]
)0LQSXW
PRGXODWLQJ
VLJQDO
IF 0+]
9&2
1
·1
·0 IF 0+]
%DVHEDQG
·1 9&2
1
0+]
4XDGUDWXUH
)0GHWHFWRU
·0
0
Problem 1.4 In the wideband receiver depicted in Figure 1.19, the first VCO is tuned
at 3.5 GHz,
(a) Find the received signal frequency.
(b) What is the first image frequency of the first mixer? Is it in the receive band? In
this situation, the image frequency in the second mixer can be emanated from two RF
components. Determine those components’ frequencies at the antenna RF input.
(c) If the input low-pass filter is a first-order one with a 3 dB corner frequency of 1 GHz
and the other two bandpass filters have a frequency response given in Figure 1.19,
determine the attenuation values of the first image and those two RF components which
could result in the second image in this circuit.
BW=10MHz BW=2MHz
1GHz fc=3GHz fc=21.4MHz
0
log(f) 3GHz
log(f) 21.4
log(f)
MHz
1GHz 10MHz
2MHz
Problem 1.5 In the WiMax receiver depicted in Figure 1.20, the input frequency
range is between 3.4 GHz and 3.6 GHz with 20 MHz channel spacing. The input
band-pass filter is of second order with the bandwidth of 400 MHz and the center
frequency of 3.5 GHz. The first IF signal is at fin /5 where fin is the input signal
frequency. The second IF frequency is zero. The output low-pass filters are of third
order with a corner frequency of 10 MHz. Moreover, the frequency response of the
amplifier A and the mixers are constant.
(a) What is the image frequency at the receiver input which experiences the maximum
attenuation, in dB?
(b) Find the division ratio M for the receiver for the input frequency range.
(c) If the input frequency is 3.5 GHz, what is the attenuation of the unwanted adjacent
and the unwanted neighboring channels? (channel spacing is about 20 MHz )
10MHz
I
A
fc=3.3-3.7GHz
A IF1
10MHz
Q
A
π/2
÷16 VCO ÷4
16MHz
÷M
Problem 1.6 In a radio receiver depicted in Figure 1.21, we have cascaded the blocks
with the given specifications.
(a) If at the input two signals (the main channel and the adjacent channel, respectively)
with a power of −60 dBm each at frequencies of 900.060 MHz and 900.120 MHz are
present, what is the power of IM3 components at the output of the mixer?
(b) If the 45 MHz band-pass filter has a passband of 60 kHz and is of second order,
then what is the output of the 45 MHz filter emanating from the IM3 component at
900 MHz. (See IM3 concept in Chapter 4.)
ȕ:1
2. Oscillators
Ga (s)
H (s) = (2.1)
1 − β Ga (s)
28 Chapter 2. Oscillators
)UHTXHQF\
1RQOLQHDU VHOHFWLRQ
DPSOLILHU
=/
*D
ȕ
'LYLGHU
The loop will be unstable (the oscillation will occur) once β Ga (s) = 1 which is called
the Barkhausen condition. Consider the noise at the input of the amplifier which is
amplified and passed through the tank circuit and is fed back through the divider to
the input of the amplifier
√ (in this example, we have the resonant frequency of the tank
which is f0 = 1/(2π LC)), and by means of a nonlinear amplifier, the returning signal
would be in phase at the input. Typical white noise samples in the frequency domain
and in the time domain are shown in Figure 2.2.
Thermal noise has a white spectrum; in other words, it has a fixed power spectral
density at least in the RF range. The power spectral density (in W/Hz) can be expressed
as N0 = kT and the RMS voltage across a resistance R can be expressed as
Vn 2 = 4 kTBR (2.2)
V W/Hz
0 t
f
0
(a) (b)
Figure 2.2: Typical white noise samples, (a) in time and (b) in frequency
domain.
2.2 First Approach: Positive Feedback 29
at a specific time. Thus, the noise level is usually specified by its power spectral
density or its RMS value. Now considering the noise shown in Figure 2.2 passes
through a frequency selective circuit, the noise spectrum will be changed according to
the resonant circuit response. This will result in rejection of the noise in frequencies
out of the passband of the resonant circuit. White noise is transformed to colored
noise with a selective skirt shaped spectral density. Its time-domain waveform would
resemble approximately a sinusoid, if the resonant circuit has a high quality factor.
Then, the shaped noise passes through the divider and the amplifier, and the amplifier
amplifies it and once again it is applied to the amplifier input after the division. If
the loop gain of the system is larger than unity, the fed back noise would build up
until the nonlinearity of the amplifier compresses the gain and the loop gain of the
system approaches unity. It should be noted that the larger the fed back signal becomes
through this process, the more it will approach a pure sinusoid. Thus, when looking
at the output, there is approximately an amplified noisy sinusoid whose spectrum
approaches a pseudo-impulse-shaped spectrum in the frequency domain. The larger
the quality factor of the frequency selective circuit, the better the spectral purity of the
output signal. That is, the output signal would approach the sinusoidal form. That is
because of better rejection of the noise sidebands by a sharper filter. This process is
shown in Figure 2.3.
As depicted in Figure 2.3, for high quality factor tank circuit, the output signal is
more similar to a sinusoid; however, degrading the quality factor will result in more
phase/amplitude noise and higher harmonics level.
Vout Vout
f t
Vout Vout
f t
Figure 2.3: The output signal of the resonance circuit for different quality factor
values.
30 Chapter 2. Oscillators
Now let’s consider the amplifier nonlinearity. We would now investigate the
definition of the gain in a large-signal/limiting regime. Table 2.1 shows the output
voltage versus the input voltage of a saturating amplifier.
The large-signal gain of the amplifier is defined as the ratio of the output fundamen-
tal to the input fundamental voltage of the circuit. As Table 2.1 suggests, increasing
the signal level will result in decrease in the large-signal gain. This limiting behavior
of this circuit will stabilize the oscillation amplitude eventually. Furthermore, charac-
teristic of a typical nonlinear amplifier is illustrated in Figure 2.4 and Figure 2.5 which
demonstrates the limiting behavior of the amplifier for large signals.
We have assumed an implicit approximation in our above descriptions. By entering
the large-signal regime due to the nonlinear behavior of the circuit, a number of
harmonic frequencies of the fundamental signal will be generated as well. We can
define a more precise definition for an effective large-signal gain of the amplifier as
the ratio of the first harmonic amplitude at the output to the fundamental input voltage
amplitude, while neglecting the other harmonics at the output.
How does a positive feedback loop for an oscillator stabilize? With respect to the
definition of the effective large-signal gain of an amplifier, the Barkhausen’s criterion
suggests that the loop should have a unity gain with zero phase at the oscillation
frequency (Ga β = 1∠0), as such the loop will be stabilized. It should be noted that
the feedback in any oscillator is positive which results in initial noise amplification
and consequent oscillation. There are a number of important parameters in oscillator
Table 2.1: Effective gain with input and output signal of a saturating amplifier.
Vo
5.2
3.5
0 Vi
0.25
0.5
design, such as topology, resonant circuit, small-signal loop gain, large-signal loop
gain, signal amplitude, and phase noise. Phase noise in the oscillators is one of the
most interesting and challenging issues. We briefly discuss about the phase noise
here. Phase noise is the result of the interaction of baseband white noise signal with
the sinusoidal oscillation signal in the nonlinear circuit of the oscillator, in the sense
that the amplitude and the phase of the sinusoidal oscillation signal are modulated by
the random noise signal. As such, we can describe the general output of a sinusoidal
oscillator as V (t) = (A + an (t)) cos (ω0t + φ0 + φn (t)), where dφ n an
dt ω0 and A 1.
Here A is the amplitude of the oscillations, an (t) represents the random amplitude
modulation, ω0 is the radian frequency of oscillation, φ0 is the phase of the oscillation,
and φn (t) represents the random phase modulation. In Figure 2.6, the oscillations’
start-up of a typical oscillator as a function of time is depicted.
Amplitude
100mV
-100mV Time(µs)
0 1 2 3 4 5 6 7 8
If one computes the auto correlation function of this signal, and takes the Fourier
transform of it, he/she will obtain the spectral density of the oscillator signal. This
signal would be in the form of a narrow skirt around the sinusoidal carrier. The
importance of the phase noise is in the coherent receivers. In the sense that one intends
to detect, for example, different phase modulating levels on the carrier signal, the
random phase noise of the carrier induces a random phase shift at the output of the
detector. This exacerbates the signal detection process. We refer the interested reader
to more advanced texts regarding sinusoidal oscillators and the phase noise for further
investigation [5].
RL
VO = VS = −10VS (2.3)
RL + RS + R1
As it is obvious, the output voltage is larger than the input voltage and excess power is
generated indeed. A similar concept is shown in Figure 2.8.
As it is obvious, there is no external excitation in the circuit in Figure 2.8 except
the noise current. The existing thermal noise in the resistor (random movement of
electrons) may be amplified in the circuit by means of positive feedback.
IL GL
= (2.4)
In GL + GS + jCω − Lωj
(QHUJ\ 5
GLVVLSDWRU
ȍ 92
(QHUJ\ 56
JHQHUDWRU
(QHUJ\
GLVVLSDWRU ȍ 5/
(QHUJ\ ȍ
JHQHUDWRU
96
IL
GL L C GS
-25mƱ 20mƱ In
In2 =4kTBGS
At the resonant frequency, the imaginary part vanishes, and consequently, the above
equation reduces to
IL −25
= =5 (2.5)
In −25 + 20
As such, the noise current will be obviously amplified. If one calculates the overall
tank, conductance will reach to the total Geq = −5 mf, and thus there will be a net
energy generator in the circuit. This negative conductance will amplify the noise and
the frequency response of the tank shapes its spectrum in a way as described in the
previous section. Once the large-signal oscillation is established, one can neglect the
noise current source and writing the KCL at the common node of the circuit, one would
obtain
YL (V )V +Y ( jω)V = 0 (2.6)
As V has a nonzero value, therefore,
-R(A) L C RS
where A is the amplitude of the current phasor. I having a nonzero value, we would
have
1
−R(A) + RL + j Lω − =0 (2.9)
Cω
As such, the oscillation condition simplifies to
ZT ( jω, A) = 0 (2.10)
where ZT ( jω, A) stands for the total loop impedance.
Solution:
No, because the oscillator will never start. We have stated that for starting the
oscillations the net negative conductance must be less than zero and the circuit by
its nonlinear behavior will decrease the net negative conductance to zero, and the
oscillation will be stabilized.
It can be asserted that most types of oscillators can be analyzed by either of the two
methods presented in the previous sections.
Z2
+
Vi AV
- +
Z1 Z3 Vo
-
In this topology, the positive feedback is realized by the voltage division through
Z1 and Z2 . Normally, the three external elements in the oscillator circuit are purely
reactive elements. Now, consider Z1 ≈ jX1 , Z2 ≈ jX2 , and Z3 ≈ jX3 . The three reactive
elements should resonate at the oscillation frequency. For the oscillation condition,
one can write
X1
Av = 1 Unity loop gain condition (2.11)
X1 + X2
X1 + X2 + X3 = 0 Resonance condition (2.12)
C2 L2
Vo Vo
Vi Vi
L3 C3
L1 C1
(a) (b)
Figure 2.11: Two possible oscillator topologies with negative voltage gain
(Av < 0).
36 Chapter 2. Oscillators
L2 C2
Vi Vo Vi Vo
L1 C3 C1 L3
(a) (b)
Figure 2.12: Two possible oscillator topologies with positive voltage gain
greater than unity (Av > 1).
/ &
9L & 9L /
& / 9R
9R
D E
Figure 2.13: Two possible oscillator topologies with positive voltage gain less
than unity (0 < Av < 1).
VCC
RFC
R1 L
L
Q
CB
Q C2 C1
C2
R2 C1
RE CE
Figure 2.14: The common-emitter oscillator circuit with 180◦ phase shift
through the LC circuit.
VCC
RFC
R1 L
Q C2
CB L
Q C2
RE C1
C1
CB R2
RE
VCC
R1 L
Q
CB C1
Q L
C1 C2
C2 RE
R2
RE
the emitter by the common-collector voltage gain. This gain is slightly less than but
near to unity and the output is in phase with the input. The output voltage is fed back
to the base by the capacitive step-up transformer. At the frequency where the positive
feedback is realized and once the loop gain is compressed to unity, the oscillations will
materialize. Note that here the transistor amplifier has a voltage gain of less than unity
but a current gain larger than unity and consequently a power gain greater than unity.
Note that in all the above oscillators, there is a resonant and dividing circuit which
consists of an inductor and two capacitors which is called the Colpitts oscillator. The
dual of these circuits could be used as an oscillator as well, that is, with a resonant
circuit of a single capacitor and two inductors which is called the Hartley oscillator.
C1 +C2
Lω − = 0 for Colpitts oscillators (2.14a)
C1C2 ω
1
Cω − = 0 for Hartley oscillators (2.14b)
(L1 + L2 ) ω
VCC VCC VCC
L
RFC RFC
R1 L R1 L R1 L
CB CB CB Q
C2 C2
Q Q Q
C2 C1 C2 C1 C1
R2 C1 R2 CB R2
RE CE
RE RE Colpitts
2.4 Oscillator Topologies
Core
(a)
L2 RFC
R1 R1 R1 C
C C C
Q
Q Q Q L2
CB CB L2 CB L2 L1
R2 L1 R2 L1 CE R2 L1 CE
CB Hartley
RE CE Core
RE RE
(b)
Figure 2.17: The Colpitts and the Hartley oscillators (in common-emitter, common-collector, and common-base configura-
tions) and their corresponding main core circuits.
39
40 Chapter 2. Oscillators
Either of these two equations show a parallel resonance condition (why?) in the
reactive elements surrounding the transistors. Furthermore, there is a positive feedback
from the output to the input (180◦ phase of the voltage gain and 180◦ phase of the
voltage division for CE case, and zero-degree phase of the voltage gain and zero-degree
phase of the voltage division for the CB and CC cases).
M;
9 &I 9R
Solution:
For determining the oscillation condition, we just absorb the parasitic elements of
the transistor into the surrounding reactances as follows
−1
1 1
Z1 = + + jCi ω (2.15)
jX1 ri
−1
1
Z2 = + jωCf (2.16)
jX2
1 −1
1
Z3 = + jωCo + (2.17)
ro jX3
As such, we can consider the following equivalent circuit for this oscillator.
Z2
+ +
Vo GmV1 Z3 Z1 V1
- -
−GmV1
Vo = 1
(2.18)
Y3 + Z2 +Z 1
−GmV1 Z1
V1 = 1
× (2.19)
Y3 + Z2 +Z1 Z1 + Z2
−Gm Z1
=1 (2.20)
1 +Y3 (Z2 + Z1 )
Or
−Gm Z1 Z3
=1 (2.21)
Z1 + Z2 + Z3
CS
CP rS
LS
f
fS fP
As it is obvious from the crystal model in Figure 2.20, there exists a series and
a parallel resonance for the crystal. That is the crystal impedance, at first, falls to a
very small value at its series resonance frequency ( fS ) and then changes to a very high
impedance at its parallel resonance frequency ( fP ). Then, its impedance due to the
parallel capacitor falls gradually to a low impedance value at frequencies much higher
than fP . The values of the equivalent circuit element of the crystal are such that the
difference between the series and the parallel resonance frequencies is quite small
while the difference between the impedances at these two frequencies is quite large.
The input impedance of the crystal can be easily computed as
1 − ω 2 LsCs + jωrsCs
Z( jω) = (2.22)
jωCp 1 +Cs /Cp − ω 2 LsCs + jωrsCs
1
Given the fact that rs ωsCp , the crystal impedance value at ωs with a very good
approximation would be
Z( jωs ) ≈ rs (2.23)
This is a quite small value for the crystal impedance. On the other hand by the fact that
Q = Lsrωs s has a very large value and CCps 1, then the crystal impedance value at ωp
with a very good approximation would be
QCs
Z( jωp ) ≈ (2.24)
ωpCp 2
This impedance value is normally quite large. Interestingly, the crystal impedance
above fs and under fp is inductive (with a large inductive reactance derivative, ∂∂ ωX ) and
its impedance for frequencies under fs is capacitive (with a large capacitive reactance
derivative, − ∂∂ ωX ). This is the phenomenon which stabilizes the oscillation frequency in
the crystal oscillators. Figure 2.21 shows the computational results of a typical 10 MHz
crystal impedance with Cs = 9.1 fF, rs = 35 Ω, Ls = 27 mH, and Cp = 2 pF. The series
resonant frequency of the crystal would be 10.153542 MHz and the parallel resonant
frequency would be 10.176615 MHz. The quality factor of the crystal would be 49200.
As such, the crystal impedance goes from 35 Ω to 1.75 MΩ within a frequency span
of 23 kHz only. At low frequencies, the capacitors are open circuit, and therefore
the impedance will be high. The first resonance frequency in the circuit is due to the
120 90
120 100
110 80 75 0
100 50
|Z|(dBΩ)
Z(Deg.)
90 40
10 10.2 10.4
25 -90
10.1 10.2 10.3
80 0
70 -25
60 -50
50 -75
40 -100
2 4 6 8 10 12 14 16 18 20 2 4 6 8 10 12 14 16 18 20
Frequency (MHz) Frequency (MHz)
resonance of the series inductor Ls and the series capacitor Cs which is called the series
resonance fs . At this frequency, the inductor and the capacitor will tune out and make
approximately a short circuit (very low impedance) at the right-hand branch. Thus,
the overall impedance of the crystal will be equal to rs k jωCp . By increasing the
frequency, Cs goes low impedance and the right-hand branch becomes inductive, the
next resonance will occur approximately due to the resonance of Ls and Cp which we
call the parallel resonance and show it by fp . Finally, if we continue increasing the
frequency, the parallel capacitor will short out the whole crystal impedance at high
frequencies. The series resonance frequency can be calculated from Equation 2.25:
1
fs = √ (2.25)
2π LsCs
The parallel resonance frequency can be calculated from Equation 2.26:
q
fp = fs 1 +Cs /Cp (2.26)
Crystal is generally used in oscillators where the acoustic vibration occurs in the body
of the crystal, and by the virtue of the piezoelectricity of the crystal, those oscillations
are transformed into electrical oscillations. A crystal has a very high quality factor
which will result in the purity of the oscillator signal spectrum. This quality factor for
crystal in Figure 2.20 is equal to 49200. As stated in Equation 2.24, the maximum
impedance of the crystal occurs at the parallel resonant frequency and it has a very
high value described by this equation.
A number of crystal oscillator topologies are shown in Figure 2.22.
In Figure 2.22(a) which is a common-collector Colpitts-like oscillator, the crystal
acts as an inductor. That is the inductive reactance of the crystal resonates with the
capacitances of C1 and C2 . The feedback voltage across RE appears through the step-up
capacitive transformer across the crystal terminals. The oscillation frequency would be
slightly above fs .
In Figure 2.22(b) which is a common-base Colpitts-like oscillator, CB is a large
(short-circuit) capacitor. The crystal is in series within the feedback loop. The resonant
D E F
frequency of the LC circuit should be approximately the same as the series resonance
frequency of the crystal. the oscillation frequency would be approximately fs .
In Figure 2.22(c) which is a common-base Colpitts-like oscillator, CB is a large
(short-circuit) capacitor. The crystal is put in parallel with the capacitive divider and it
resonates with these capacitors at the oscillation frequency. The inductive reactance of
the crystal is tuned out by the capacitors. The oscillation frequency would be slightly
above fs .
Normally, a piece of quartz crystal has several electro-acoustic resonant modes.
These modes are called overtones which occur approximately at the odd multiples of
the fundamental resonant frequency. A more generalized circuit model of a crystal is
shown in Figure 2.22. In this model, the higher order resonances (overtones) of the
crystal are shown by the additional parallel RLC branches in the circuit. Normally, the
higher order modes resonances have a lower Q than the fundamental mode resonance.
As depicted in Figure 2.23, a crystal may have a number of higher order resonant
frequencies. Thus, by choosing the main oscillation frequency in the crystal circuit
meticulously, one may use it for higher desired overtone. For instance, in Figure 2.22(c)
where there is no inductor, the circuit is forced to oscillate at the fundamental fre-
quency of the crystal. That is to say the crystal will become purely inductive near
the fundamental resonant frequency. It is possible to design a frequency selection
circuit whose frequency is a multiple of the fundamental frequency of the crystal.
For example, if one designs an LC tank with 75 MHz resonance frequency, with the
fundamental frequency of 15 MHz of the crystal, the oscillator finally will oscillate at
75 MHz. However, the price of higher oscillation frequency is injecting more energy
and as a result more power loss and therefore lower quality factor. In reality, one may
order the manufacturer to make a crystal with a specific parallel resonance frequency
by realizing specific parallel capacitor. As a rule of thumb, a parallel capacitor is
usually near a few pFs or a few tenths of pFs (this value can be adjusted by the crystal
manufacturer to have a precise oscillation frequency using the crystal). Moreover, the
resonance frequency variations of a crystal with temperature for a range of 0 − 70◦
Celsius is just about 50 ppm. In other words, if the oscillation frequency is 1 MHz,
by those temperature variations, the oscillator may have about 50 Hz frequency drift.
Answer:
No, because there is a technical manufacturing problem in the design of very high
frequency crystals. As a matter of fact, the crystal disc will become too thin to
fabricate, the package capacitance would increase, microphonic issues would arise,
and unwanted frequency modulation might happen. (Microphonic effect is an
acoustic frequency modulation effect once the crystal is physically shaken).
Now, assume that we have a transmitter with a carrier frequency of 900 MHz. With
50 ppm frequency variations in the reference crystal, there will be 45 kHz frequency
variation at the carrier, which in GSM with 200 kHz channel bandwidth is too much.
Thus, we need a more precise frequency for this kind of application. A temperature-
compensated crystal oscillator (TCXO) might be a solution, where by microtuning
and using temperature-dependent biasing, the temperature variations of the oscillation
frequency is compensated. Therefore, the frequency will be stabilized within a range
of few ppm’s, e.g., a frequency variation of 3 ppm here will result in a variation of
2.7 kHz in the carrier frequency. If this range of variation is not yet acceptable, a
PLL-based synthesizer carrier might be employed. We discuss more about this subject
in the next chapter.
70
65
ESR(dBΩ)
60
55
50
45
40
35
30
25
1 10 100
Frequency (MHz)
with a resonance frequency in the range of few MHz have a relatively higher series
resistance (about few hundreds of ohms to few thousands ohms), and the crystals with
a resonance frequency of few hundred MHz have a series resistance of the order of a
hundred ohms.
VCC
L
R1
Cbc
Q C2
Ccs
CB R2 Cbe RE C1
where ω0 is the resonant frequency of the LC tank, and Q is the quality factor of the
circuit. Q is a parameter which describes the ratio of the energy stored per cycle to the
power dissipation, i.e., higher Q means lower power dissipation compared to the stored
energy. The magnitude of the fraction in Equation 2.29 becomes equal to unity at the
resonance frequency ω0 . Moreover, in the magnitude response of the filter, the same
parameter Q shows the sharpness of the frequency response near the center frequency.
Now consider the frequency response of the band-pass filter as shown in Figure 2.26.
By computing the upper and the lower frequency 3 dB points in the frequency
response of the filter, it can be shown that the quality factor for Figure 2.26 can be
written as
ω0 f0
Q= = (2.31)
BW ∆f
Here, the bandwidth is described in radian frequency where BW = 2π∆ f . Here, Q is
the quality factor of the bandpass filter.
|H(jω)|
ω0-ω0/2Q ω0+ω0/2Q
ω0 ω
BW
The quality factor can also be defined for lossy elements such as inductors and
capacitors. For a lossless inductor, Q will be infinite; however, in reality, due to
different sources of loss in the inductors (series resistance, skin effect, and magnetic
core loss), they will have a finite Q. An inductor and its equivalent circuit are shown in
Figure 2.27. The quality factor of an inductor is defined as in 2.32:
1 2
2 LI Lω Rp
Q = 2π 1 2
= = (2.32)
2 rs I T
rs Lω
Note that for a specified inductor, the values of rs and Rp are quite different. As a
rule of thumb, the quality factor of a discrete inductor is between 50 and 100 and for
an on-chip inductor due to its two-dimensional structure is about 3 to 5. The other
reactive lossy element in circuits is a capacitor. For an ideal capacitor, the quality
factor is infinite. Nonetheless, for a real capacitor due to dielectric losses or its series
resistance, the quality factor is finite. The quality factor for a capacitor can be given by
Equation 2.33:
1 2
2 CV 1
Q = 2π 2 = RpCω = (2.33)
1V rsCω
2 Rp T
Note that for a specified capacitor, the values of rs and Rp are also quite different. The
equivalent circuit of a capacitor is shown in Figure 2.28.
As a rule of thumb, for a discrete capacitor, the quality factor is between 50 and
200 and for integrated capacitors, this value is roughly between 50 and 100. The
parameter Q first defined as the ratio of stored energy per cycle to the dissipated power
/
/ 5S /
UV
&
& 5S &
UV
is simply related to the resistive loss in the inductors and the capacitors. As such, a
resonator realized by a pair of elements like an inductor and a capacitor will have
a quality factor which will be less than the quality factor of either of the elements.
Therefore, realizing an LC filter with quality factors in excess of one hundred will not
be possible.
Normally, the skin effect increases the quality factor of the inductors with increas-
ing frequency, but the Q factor is reduced as the frequency passes a certain maximum
value. As a matter of fact, the series resistance of inductors increases by the square root
of frequency due to the skin effect. In practice, in discrete implementations, multiple
inductors are placed in parallel to mitigate the skin effect, and therefore, achieve
a better quality factor. The difficulty of placing a band-pass filter at the receiver’s
front-end is more clear now. Because of the low quality factor of the passive reactive
elements, high Q band-pass circuits are barely realizable in the receiver sections. This
is the reason for which to have very sharp filters with high quality factors, i.e., normally
ceramic filters or crystal filters are used in the receiver chain.
where gm is the transistor’s transconductance and rin is the base dynamic resistance as
described in Equation 2.35.
KT KT Vt
rin = β = (β + 1) = (β + 1) (2.35)
qIC qIE IE
Now, if the collector load resistance is 1 kΩ, for the collector bias current of 1 mA, the
gain of the amplifier becomes AV = −gm RL = −38.5. The derived equation is merely
VCC=5V
valid for the linear behavior of an amplifier; however, when a large signal is imposed
at the input, the gain equation should be modified. The transfer characteristics of the
bipolar transistor can be assumed as in Equation 2.36:
qvBE
ie = IES e kT (2.36)
For instance, with this exponential I −V characteristic, with an increase of one thermal
voltage ( kT
q ≈ 26 mV) at the input, the output current will be multiplied by a Neper.
With a supply voltage of 5 V in Figure 2.29, the corresponding output voltage of the
amplifier for different input levels is depicted in Figure 2.30. Here, β is assumed to be
equal to 100.
As it is obvious from Figure 2.29, the output DC of the collector is 4 V. First,
consider the input amplitude is 12.5 mV. Thus, output signal swing will be about 1 V.
If we continue increasing the input voltage to 25 mV, the signal will be limited from
the top to supply voltage and from the bottom to 2.4 V. Further entering large-signal
input regime will result in limiting the signal from the top to 5 V and from the bottom
to about 2.2 V while having a significant distortion with respect to the sinusoidal form.
Consider Figure 2.31.
5
VS=12.5mV
4 VS=25mV
VC(v)
VS=50mV
3
2
1
0 2 4 6 8 10
Time (µs)
Figure 2.30: The collector voltage, for different values of input voltage, as a
function of time.
VCC =5V
RB RL=1kΩ
C1 Vc Vo
Q C2
RS
50Ω
+
Vs IBias CBypass
Vout
= −Gm RL (2.37)
Vs
5
4
3
VC(v)
2
1
0
0 2 4 6 8 10
Time (µs)
Figure 2.32: Output collector voltage with 100 mV input at 1 MHz frequency.
9&&
/ 5/ &
5%
9RXW
4
9V
4 4
&%\SDVV
9((
9&Y
7LPHV
excos(ω0 (t)) = I0 (x)+2I1 (x)cos (ω0 (t))+2I2 (x)cos (2ω0 (t))+· · ·+2In (x)cos (nω0 (t))
(2.38)
Here the functions In (x) are called the modified Bessel functions of the first kind. These
functions are the solutions of the Bessel’s differential equation at certain conditions.
The evolution of these functions with respect to their argument has generally an
exponentially increasing form. Furthermore, the higher order function is generally
smaller than the lower order function for the same argument. That is
In+1 (x)
<1 for all x (2.39)
In (x)
Additionally
In+1 (x)
lim =1 (2.40)
x→∞ In (x)
and
I1 (x)
lim =1 (2.41)
x→∞ I0 (x)
2.10 Large-Signal Transconductance and Harmonic Tuned Amplifiers 53
,Q[ ,Q[
,[
,[
,[
,[
,[ ,[ ,[
,[ ,[
,[
,[
,[
[ [
Figure 2.35: Typical variations of the modified Bessel functions with respect to
their arguments.
Table 2.3: Numerical values of the first four modified Bessel functions of the
first kind as a function of their argument.
x 0 1 2 3 4 5 6 7 8 9
I0 (x) 1 1.27 2.28 4.88 11.30 27.24 67.23 168.59 427.56 1093.59
I1 (x) 0 0.57 1.59 3.95 9.76 24.34 61.34 156.04 399.87 1030.91
I2 (x) 0 0.14 0.69 2.25 6.42 17.51 46.79 124.01 327.60 864.50
I3 (x) 0 0.02 0.21 0.96 3.34 10.33 30.15 85.18 236.08 646.69
and
I1 (x) x
lim = (2.42)
x→0 I0 (x) 2
Furthermore, I0 (0) = 1 and In (0) = 0 for n > 1. Figure 2.35 shows the variations
of the modified Bessel functions of different orders with respect to their argument.
Table 2.3 shows the numerical values of the modified Bessel functions of different
orders as a function of their arguments.
The above equation is valid for the small-signal behavior of an amplifier. One may
write a general equation for large-signal bipolar transistor current as 2.45:
If we define x = Vi /Vt and expanding in terms of modified Bessel functions of the first
kind, that will result in Equation 2.47:
VBE
bias
iC = αIES e Vt (I0 (x) + 2I1 (x) cos (ω0t) + 2I2 (x) cos (2ω0t) + · · · ) (2.47)
While employing the current source, the DC component of current will be constant
(how?). When x increases, I0 (x) will increase similarly but VBEbias will decrease indeed.
The only mechanism that maintains the DC constant is VBE depreciation as stated. The
output signal for a large-signal input can be written as
2I1 (x)
Vout = VCC − αIEbias ZL (0) + ZL ( jω0 ) cos (ω0t) +
I0 (x)
2I2 (x)
ZL (2 jω0 ) cos (2ω0t) + · · · (2.50)
I0 (x)
where the input signal is VS = Vi cos(ωt) and α ≈ 1. The typical frequency response
of the amplifier is shown in Figure 2.36. Here, it is assumed that the output is tuned to
the first harmonic of the input.
As it is obvious from Figure 2.36, the output current contains all harmonics of
the input signal. However, by tuning the band-pass filter, any output harmonic can
be selected at the output. In tuned amplifier and oscillator applications, it is normally
assumed that the output circuit is tuned to the fundamental harmonic of the input.
While band-pass filters are of interest in narrowband applications, we investigate their
impedance behavior a little bit more. A simple parallel RLC band-pass filter fed by a
current source is shown in Figure 2.37.
2.10 Large-Signal Transconductance and Harmonic Tuned Amplifiers 55
_9RMȦ_ )UHTXHQF\UHVSRQVHRI
EDQGSDVVILOWHU
2XWSXWKDUPRQLF
Ȧ
Ȧ Ȧ Ȧ Ȧ
Figure 2.36: Harmonics of output and selecting behavior of the band-pass filter.
IS C R1 L
Zin
Figure 2.37: Resonant circuit.
RT
Zin = (2.51)
1 + jQ( ωω0 − ωω0 )
where
1
ω0 = √ RT = R1 k RPC k RPL (2.52)
LC
where RPC and RPL are the equivalent parallel loss resistances of the capacitor and the
inductor, respectively. Q is the overall quality factor of the circuit and is expressed as
2.53:
R1 k RPC k RPL
Q= = RTCω0 (2.53)
Lω0
1 1 1 Lω0
= + + (2.54)
Q QL QC R1
56 Chapter 2. Oscillators
The important point in Equation 2.54 is the dominance of low Q element which is
usually an inductor. Typical frequency response of the band-pass filter is shown in
Figure 2.38.
It can be shown that the input impedance at −3 dB point of the circuit is as 2.55:
RT
Zin = (2.55)
1 + j(1)
Equation 2.58 is of great importance. It suggests that we can calculate the quality
factor of resonant circuits by finding the ratio of center frequency to its 3 dB bandwidth.
Moreover, one may obtain the bandwidth of the circuit by the division of the resonant
frequency by the quality factor.
For the nth harmonic of the input, the load impedance described in Equation 2.57
can be expressed as
RT nRT
ZL ( jnω0 ) = 1
' 2 − 1)
(2.59)
1 + jQ n − n jQ (n
|Z(jω)|
BW
RT
-3dB
With respect to the aforementioned derivation of Q, one may obtain output voltage
of the circuit depicted in Figure 2.33 as
2RT I1 (x) 4RT I2 (x) π
Vout = VCC − αIEbias cos (ω0t) + cos 2ω0t −
I0 (x) 3QI0 (x) 2
3RT I3 (x) π
+ cos 3ω0t − +··· (2.60)
4QI0 (x) 2
Neglecting the smaller valued harmonic terms, we can simplify the above equation to
the following
2RT I1 (x)
Vout ≈ VCC − αIEbias cos (ω0t) (2.61)
I0 (x)
Or
2RT I1 (x)
Vout ≈ VCC − xαIEbias cos (ω0t) (2.62)
xI0 (x)
Or
2I1 (x)
Vout ≈ VCC − gm Vi RT cos (ω0t) (2.63)
xI0 (x)
Here we can define the large signal transconductance as the following
IC1 2I1 (x)
Gm = = gm (2.64)
Vi xI0 (x)
In the above-mentioned equation, it is assumed that the load impedance is tuned to the
first harmonic of the input. As such, it is observed that the higher order harmonics
amplitudes are decreasing monotonically as a function of amplitude and frequency.
If the load quality factor Q is sufficiently large, the higher order harmonics could be
neglected compared to the fundamental harmonic.
Furthermore, regarding Equation 2.57, once the load is tuned to the mth harmonic
of the input, the output voltage will become
mRT 2I1 (x) π 2mRT 2I2 (x) π
Vout =VCC −αIEbias × cos ω0 t+ + × cos 2ω 0 t+ +· · ·
Q (m2 − 1) I0 (x) 2 Q (m2 − 4) I0 (x) 2
2RT Im (x) nmRT 2In (x) π
+ cos (mω0 t) + · · · + × cos nω0 t − +··· (2.65)
I0 (x) Q (n2 − m2 ) I0 (x) 2
Note that, in Equation 2.65, in the developed series n 6= m and here it is assumed that
m > 2. Neglecting the smaller-valued terms, we can approximate the above equation
by the following
2RT Im (x)
Vout ≈ VCC − αIEbias cos (mω0t) (2.66)
I0 (x)
To clarify more what is described in Equation 2.65, two special cases are considered in
the following sections.
58 Chapter 2. Oscillators
2.10.1 Case I: Resonant circuit is tuned to the first harmonic of the input
frequency (tuned amplifier case)
For this special case, as it was already derived in Equation 2.60 for the first three
harmonics, in other words for n = 1, 2, 3. We define nth harmonic at the output as
Hn (x), we will have
2I1 (x)
H1 (x) = RT IEbias (2.67a)
I0 (x)
2I2 (x) RT 2I2 (x) 2RT
H2 (x) = IEbias ≈ IE (2.67b)
I0 (x) 1 + jQ(2 − 2 )
1 I0 (x) 3Q bias
2I3 (x) RT 2I3 (x) 3RT
H3 (x) = IE ≈ IE (2.67c)
I0 (x) 1 + jQ(3 − 13 ) bias I0 (x) 8Q bias
2.10.2 Case II: Resonant circuit is tuned to the second harmonic of the input
frequency (frequency multiplier case)
Here, m = 2. For this case, we derive the equations for first to third output harmonics,
in other words for n = 1, 2, 3. The equation for the first three harmonics can be written
as
2I1 (x) RT 2I1 (x) 2RT
H1 (x) = IEbias ≈ IE (2.68a)
1
I0 (x) 1 + jQ( 2 − 2)
I0 (x) 3Q bias
2I2 (x)
H2 (x) = RT IEbias (2.68b)
I0 (x)
2I3 (x) RT 2I3 (x) 6RT
H3 (x) = IE ≈ IE (2.68c)
3 2
I0 (x) 1 + jQ( 2 − 3 ) bias
I0 (x) 5Q bias
Typical output harmonic currents and the load impedance variations are shown in
Figure 2.39 for both cases (I and II).
As it is obvious from Figure 2.39, in the first case, the output band-pass filter is
tuned to the first harmonic of the input which means it attenuates higher harmonics. In
the second case, the output band-pass filter passes the second harmonic and attenuates
other frequency components, i.e., the first harmonic, the third harmonic and the higher
ones at the output. For the ease of calculation, we define a large-signal transconduc-
tance (Gm ) based on the harmonic number of the output. Our goal is to obtain an
equation for the output signal in terms of Gm . For instance, one may write the output
voltage for the first harmonic as
We may write the amplitude of input signal as Vi = xVt ; so, we derive Gm1 as
I1 I1
Gm1 = = (2.70)
Vi xVt
2.10 Large-Signal Transconductance and Harmonic Tuned Amplifiers 59
|VO(jω)|
Case Case
1 2
ω
ω1 2ω1 3ω1 4ω1
Figure 2.39: Typical output harmonics as well as load impedance variations for
cases I and II.
With respect to the above definition, we can write the expression for Gm1 as
It is possible to generalize Equation 2.72 for the nth harmonic of the input frequency
when the output band-pass filter is tuned at the mth harmonic of the input. In this case,
the ratio of the output nth harmonic to the input phasor can be described as follows
Gmn(x)
gm
1
0.8 2I1(x)/xI0(x)
2I2(x)/xI0(x)
0.6
2I3(x)/xI0(x)
0.4
0.2
0
0 1 2 3 4 5 6 7 8 9 10
x
Figure 2.40: Ratio of large-signal transconductance normalized to small-signal
transconductance as well as the conversion transconductances for the second
and the third harmonics.
x 0 1 2 3 4 5 6 7 8 9
2I1 (x)
1 0.893 0.698 0.540 0.432 0.357 0.304 0.264 0.234 0.209
xI0 (x)
2I2 (x)
0 0.214 0.302 0.307 0.284 0.257 0.232 0.210 0.192 0.176
xI0 (x)
2I3 (x)
0 0.035 0.093 0.131 0.148 0.152 0.149 0.144 0.138 0.131
xI0 (x)
Example 2.4 Consider Figure 2.33 where the input is VS = Vi cos ω0t with
ω0 = 2π (50 MHz). First, assume the band-pass filter is tuned to 50 MHz and
then assume it is tuned to 150 MHz. Derive the relations for the output voltage at
the first and the third harmonics.
Solution:
Using Equation 2.73, one may reach to Table 2.5.
2.10 Large-Signal Transconductance and Harmonic Tuned Amplifiers 61
Table 2.5: Normalized values of the first and the third harmonic voltages for
the output tuned to either of the first or the third harmonics.
H1 (x) , n = 1 H3 (x) , n = 3
Vo1 2I (x) Vo3 2I (x) RL
Vi = gmQ xI01 (x) RL cos 108 πt 8
Vi = gmQ xI03 (x) 1+ jQ 8 cos 3 × 10 πt
3
Vo1 2I1 (x) RL 8 Vo3 2I3 (x)
Vi = gmQ xI0 (x) RL cos 3 × 108 πt
Vi = gmQ xI0 (x) 1− jQ 8 cos 10 πt
3
Apparently, in the first case, we have first harmonic at the output with a high
gain and third harmonic at the output with a lower gain. In the second case, we
have the first harmonic at the output with a low gain and the third harmonic at the
output with a relatively higher gain.
VCC
R
C R L 50Ω
R
RB 50Ω
Vout
Q1
+
Vs
Q3 Q2
CBypass
-VEE
In the next section, we focus on oscillators based on tapped capacitor and tapped
inductor transformers. The inductive transformers are tunable with their number of
turns and have a good isolation.
VCC
R1
+ Q CE M12
CB V
R2 1 -
RE L2 L1 RL C1
VCC
Figure 2.42: A common-base tuned circuit oscillator.
Solution:
The equivalent circuit for the above oscillator is depicted in Figure 2.43.
M12
1:
L1
- -
GmV1 RL C1 L1 Vo Gin GE V1
+ +
Figure 2.43: The equivalent circuit for the common-base tuned circuit
oscillator.
GmV1
Vo = 2 (2.74)
1
RL + jC1 ω − L1jω + ML12
1
(Gin + GE )
M12
L1 G m
2 =1 (2.76)
1
RL + jC1 ω − L1jω + M12
L1
Gm
α + GE
Separating the real and the imaginary parts of Equation 2.76, one obtains two
distinct equations
1
C1 ω − =0 (2.77)
L1 ω
2
1 M12 1
RL + L RE
Gm (x) = 1 (2.78)
M12 1 M12
L1 1 − α L1
From the two above equations, the first one gives the oscillation frequency and the
second one through Gm (x) would determine the oscillation amplitude.
v
ic2 = αIES eq(VBE0 +v2 )/kT = αIES eq(VBE0 − 2 )/kT (2.80)
VCC
LL RL CL CL RL LL
IC1 IC2
Vout
+
Q1 Q2
+
v
IE
VEE
IC IC h z i
ic2 = = 1 − tanh (2.84)
1 + e+z 2 2
where
qv v
z= = (2.85)
kT Vt
Assuming a large sinusoidal input voltage as
v = V1 cos (ωt) (2.86)
Either of the collector AC currents becomes
IC x
ic1,2 = ± tanh cos (ωt) (2.87)
2 2
where
qV1 V1
x= = (2.88)
kT Vt
Now using the above equations, one can compute the harmonic components of the
collector currents as
1 π 1 x
Z
an (x) = tanh cos (θ ) cos (nθ ) dθ (2.89)
π −π 2 2
Note that given the fact that the differential pair transfer characteristic has an odd
symmetry, an (x) functions would be zero for even values of n. Using the fundamen-
tal harmonic current of either of the collectors, one can calculate the large-signal
transconductance of the differential bipolar stage:
IC1 IC2 qIC a1 (x) 4a1 (x)
Gm (x) = =− = = gm (2.90)
V1 V1 kT x x
where
∂ ic1 ∂ ic2 qIC IC
gm = =− = = (2.91)
∂v ∂v 4kT 4Vt
Note that the DC and the fundamental harmonic output voltages at either of the
collectors become
2.12 Inductive and Capacitive Dividers (Impedance Transformers) 65
1
0.9
0.8
Gm(x)/gm
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
10
12
14
16
18
20
0
8
Vin/VT
−
Vout = VCC − Gm (x) RLV1 cos (ωt) (2.92)
+
Vout = VCC + Gm (x) RLV1 cos (ωt) (2.93)
Using this large-signal transconductance, one can compute the amplitude of oscillations
in a differential pair oscillator.
L V
Vs 1 VL 1
Rin = = m = 2 = 2 RL (2.95)
is miL m iL m
For example, if the load resistance is 1 kΩ, for m = 5, the input impedance will be 40 Ω.
Transformers play a crucial role in communication circuits for matching purposes.
As stated earlier by tuning the resonant band-pass filter, one may attain a desired
harmonic of the input signal at the output of a nonlinear circuit. The main role of
transformers is to extract the desired signal from the resonant circuit without degrading
its quality factor. However, it is instructive to know more about inductors before
introducing their use in transformers. Discrete inductors have a good quality factor just
for low-frequency applications. These inductors will fail at high frequencies due to
their parasitic capacitances and resistances. Nowadays, RF engineers desire to integrate
every thing on a single chip, thus on-chip inductors are of interest. However, because
66 Chapter 2. Oscillators
of their planar implementation, they will not have a good quality factor. Moreover,
these elements are relatively huge and bulky in size and their fabrication occupies a
huge area on the RF chip. It is possible to implement inductors at high frequencies
(say at a few GHz) using microstrip or printed circuit transmission lines as well.
Furthermore, realization of printed inductors is possible for monolithic microwave
integrated circuits at frequencies well above 5 GHz. In the next section, we introduce
circuits for impendence transformation and step-up and step-down voltage concepts.
RP
Qunloaded = (2.96)
Lω
is 1:m is/m
+
+
Vs RL VL=mVs
Rin
C RP
Figure 2.47: Inductive transformer with its loss and a capacitive load.
2.12 Inductive and Capacitive Dividers (Impedance Transformers) 67
(TXLYDOHQW
LQGXFWRU
P
/ & 53
,GHDO
WUDQVIRUPHU
RS=50Ω
1:m
+
Vs L C RP
RP || RS × m2
QLoaded = (2.97)
Lω
Thus, as per Equation 2.96, the loading of the resonant circuit by the low impedance
source degrades the quality factor of the resonant circuit. However, if the value of
m2 RS is much larger than RP , the loaded quality factor would not be degraded as much.
Capacitive impedance transformers act like the inductive ones except that they have ca-
pacitive reactances instead of inductive ones at their input. Note that the modeling and
the behavior of capacitive/inductive coupling circuit are a straightforward procedure
and we focus here on their basic behavior.
Figure 2.50 shows a capacitive impedance transformer.
68 Chapter 2. Oscillators
At the resonance, one may obtain the input impedance of Figure 2.50 as
R
Rin = (2.98)
m2
where
C1 +C2
m= (2.99)
C1
Here, m > 1 means a step-up impedance transformation. Moreover, it is also possible
to define an equivalent capacitance as
C1C2
Ctotal = (2.100)
C1 +C2
Provided that for Rin (C1 +C2 ) ω ≥ 10, the above circuit can be modeled by a parallel
RLC circuit along with a step-up transformer as depicted in Figure 2.51. Furthermore,
the resonance condition becomes
C1 +C2
Lω − =0 (2.101)
C1C2 ω
The quality factor of this circuit will be
C1C2
Q= ωRT (2.102)
C1 +C2
where RT is the total equivalent parallel resistance of the tuned circuit. For large m, the
relative values of the capacitors usually follow CC21 1. For C2 C1 , Equation 2.102
can be approximately written as
Q ≈ C1 ωRT (2.103)
For instance, if C1 = 10 pF and C2 = 70 pF, then m = 8 and Ctotal = 8.75 pF, and a load
resistance of 1000 Ω will be transformed to 15.625 Ω. The equivalent circuit of this
step-up resonant impedance transformer is depicted in Figure 2.51 along with its dual
counter part which consists of an inductive step-up impedance transformer.
C1
L R
C2
Rin
1:m
L1 L1+L2
C GL C
GL L2
Gin=GL/m2
1:m
C1C2
C1
C1+C2
L GL L
GL C2
Gin=GL/m2
Before using the models of Figure 2.51 for the step-up transformers, let’s verify
Equation 2.98 through 2.103 using the admittance or the impedance matrices of the
corresponding circuits. Now, let’s consider Figure 2.52.
One may obtain the admittance matrix of Figure 2.52 as
I1 C1 +C2 −C1 V1
= jω (2.104)
I2 −C1 C1 V2
Or in expanded form
Now we derive the relation for the output admittance when the input is loaded with the
conductance, GL . The circuit is shown in Figure 2.53.
We can write at the input
I1 = −GLV1 (2.106)
I2
I1 C1 +
+ V2
V1 C2
I2
+
I1 C1
+ V2
V1 GL C2
Yout
Finally, we have
V1 jωC1
= =A (2.109)
V2 GL + jω (C1 +C2 )
Using Equation 2.105b, one obtains
Now, considering a high quality factor for this circuit, it is imposed that the short-circuit
quality factor should be larger than unity
Therefore, considering the fact that (C1 +C2 )2 ω 2 G2L , Equation 2.113 is reduced to
Equation 2.116:
2
C1 C1C2 GL
Yout ≈ GL + jω = 2 + jωCtotal (2.116)
C1 +C2 C1 +C2 m
C1 +C2
where m = C1 . Finally, the overall quality factor will be
If (C1 +C2 ) ω > GL and C2 > C1 , and the quality factor is greater than 10, the
equivalent circuit shown in Figure 2.51 will be valid. However, if the quality factor is
less than 10, the precise relation for the calculation of the admittance (Equation 2.113)
should be used.
The same method applies for inductive transformers. Consider Figure 2.54.
Impedance parameters of Figure 2.54 can be derived as
V1 L L2 I1
= jω 2 (2.118)
V2 L2 L1 + L2 I2
Or in expanded form
I1 = −GLV1 (2.120)
I2
I1 L1 +
+ V2
V1 L2
I2
+
I1 L1
+ V2
V1 GL L2
Yout
I2 1 1 + jωL2 GL
Yout = = jωL2
= (2.124)
V2 jωL1 + 1+ jωL jωL 2 + jωL1 − ω 2 L1 L2 GL
2 GL
Given the fact that, in the tapped inductor transformer, the short-circuit quality factor
should be greater than unity, it is deduced that
L1 + L2
> GL (2.125)
L1 L2 ω
As stated earlier in capacitive step-up transformer, with the condition stated in rela-
tion 2.125, if the quality factor of the inductive transformer is greater than 10 and
L1 > L2 , the above approximation will be valid, and the equivalent circuit shown
in Figure 2.51 can be used. Otherwise, the precise relation for the admittance
(Equation 2.124) should be employed.
2.12 Inductive and Capacitive Dividers (Impedance Transformers) 73
C1
R C2 Rin
Solution:
(a) Here the short-circuit quality factor becomes
(C1 +C2 ) ω
= 10 (2.127)
GL
As it is obvious from Equations 2.128 and 2.129, the results are quite close to each
other.
(b) Here the short circuit quality factor becomes
(C1 +C2 ) ω
=1 (2.130)
GL
74 Chapter 2. Oscillators
As it is obvious from Equations 2.131 and 2.132, the results don’t match completely,
that is, the approximation is not valid for a low quality factor circuit.
VCC
L
Q C1
RP
C2
RB
-VCC
+
Q C1
LP RP
Vout
+
C2 V
Is i
- -
Q
*P9L &HT /3 53
9RXW *LQ 9L
Here, we have
C1 C1C2
n= , Ceq = (2.134)
C1 +C2 C1 +C2
The input emitter voltage Vi is (note that here m < 1)
Vi = nVout (2.135)
Now, replacing Vout in Equation 2.133, we arrive at an expression for the closed-loop
gain of the oscillator:
nGm
ACL ( jω) = 1
=1 (2.136)
jωCeq + + R1P + n2 Gin
jωLP
As it is seen in Equation 2.136, the right-hand side of the equation is purely real, so
we can separate the real and the imaginary parts of Equation 2.136 and obtain the
following pair of equations:
1
ωCeq − =0 (2.138)
ωLP
1
RP
Gm = (2.139)
n 1 − αn
Equation 2.138 gives the oscillation frequency and by Equation 2.139, we can obtain
the oscillation amplitude through the large-signal transconductance. Here our focus
was on the first harmonic, because other harmonics are attenuated by the high-Q tuned
resonant circuit to some extent.
Example 2.8 Consider Figure 2.60, where the transistor has a current gain
α = 0.99 and a parasitic collector–base capacitance of 0.2 pF and a parasitic base–
emitter capacitance of 5 pF. Given the 1.5 mA emitter current source, compute the
oscillation amplitude and the oscillation frequency in this circuit.
S)
4 S)
Q+ Nȍ
Solution:
An important point in this example is the absorption of parasitic capacitances in
the resonant circuit. First, we calculate the capacitive transformer ratio n as
10
n= = 0.087 (2.140)
10 + 100 + 5
From Equation 2.139, we obtain
1
1000
Gm 1 = = 12.6 mf (2.141)
0.087 1 − 0.087
0.99
1.5 mA
gmQ = 0.99 × = 57.1 mf (2.142)
26 mV
Thus, using Equations 2.141 and 2.142, we will have
Gm 1
= 0.220 (2.143)
gmQ
The voltage obtained from Equation 2.144 is that of the base–emitter junction. The
collector voltage is higher by the ratio of 1/m, thus
x 0.221
VC = = = 2.54 V (2.145)
n 0.087
By assuming the supply voltage equal to 5 V, the output voltage will be
Note that the parasitic collector–base capacitance will be added to the equivalent
capacitance of the capacitive divider. So, the total capacitance would become
C1 (C2 +CBE )
Ctotal = +CBC = 9.33 pF (2.147)
C1 +C2 +CBE
1 1
f0 = √ = 164.77 MHz (2.148)
2π 100 nH × 9.33 pF
Here we verify the required condition for the capacitor tapped transformer model
that is
(C1 +C2 ) ω
= 114 (2.149)
GL
which is much larger than unity and therefore, the tapped transformer equivalent
circuit is valid here.
2.13.1 Increasing the Quality Factor and the Frequency Stability with a Crystal
Consider Figure 2.61.
As it is depicted in Figure 2.61, for increasing the quality factor, a crystal is placed
in series within the feedback loop. Normally, the parallel RLC circuit’s resonance
frequency is chosen the same as that of the crystal. However, if the frequency of
78 Chapter 2. Oscillators
Cµ
Q C1
L RL
Cπ IE C2
-VCC +VCC
resonant circuit is slightly different from the crystal series resonance, the oscillation
frequency will change to satisfy the Barkhausen’s oscillation condition. Any phase
change in the loop should be compensated by the crystal. However, due to sharp phase
characteristic of the crystal, this will result in a very small frequency change. It is
possible to show the series crystal equivalent model as presented in Figure 2.62.
Now, we are going to investigate the resonant circuit detuning more precisely. This
will result in Q degradation. Figure 2.63 shows the crystal impedance behavior about
its series resonance.
Similarly, the impedance of the parallel resonant circuit is shown in
Figure 2.64.
In this case, assume that the resonant frequency of parallel RLC is higher than the
crystal resonant frequency. The circuit will oscillate near the crystal resonant frequency
( fS ) due to its higher quality factor. The tank circuit introduces a finite phase change
as ∆Φ in the loop gain. Thus, the crystal phase characteristics should compensate this
phase by introducing −∆Φ in the loop gain to maintain the Barkhausen’s oscillation
condition. While the phase characteristics of the crystal is sharp, this phase change
does not alter the oscillation frequency significantly. It is noteworthy that the amplitude
of oscillation might be altered a little bit as well. The interested reader is referred to
section 2.20 for further details.
In another topology, the crystal might be used as an inductive reactance within the
resonant circuit of a Colpitts oscillator as depicted in Figure 2.65.
Here, with a slight shift of the oscillation frequency with respect to the crystals’
resonant frequency, the Barkhausen oscillation condition can be satisfied. Assuming
the oscillation frequency near to fs , and further neglecting the effect of CP , the crystal
impedance can be represented by
ZX = R + jX (2.150)
LS CS RS
ZX
|ZX|
rs
ZX fS f
90
fS f
-90
Figure 2.63: Impedance behavior of the crystal about its series resonance.
|Ztank|
RP
Ztank fS ftank f
90
∆Φ
fS ftank f
-90
Figure 2.64: Variations of the impedance of the parallel resonant circuit about
the resonance frequency.
where
R = rs (2.151a)
∆f
X = 2Q0 rs (2.151b)
fs
Here rs is the crystal’s series resistance and Q0 = Lsrωs s is the crystal’s unloaded quality
factor. For a series resonant circuit, one can write
1
ZX = rs + jLs ω − j (2.152)
Cω
80 Chapter 2. Oscillators
-VCC
RFC
Q C1
IE C2
-VCC
Or
ω ωs ∆ω
ZX = rs 1 + jQ0 − = rs 1 + j2Q0 (2.153)
ωs ω ωs
Here, ω = ωs + ∆ω. For the resonance condition, we should have
C1 +C2
X= (2.154)
2π fsC1C2
or
C1 +C2 ∆f
= 2Q0 rs (2.155)
2π fsC1C2 fs
The point is that we have put fs instead of fo in the left-hand side of Equation 2.155.
The reason being the fact that ∆ f is extremely small compared to fs . By resolving
Equation 2.155, one simply obtains ∆ f and the oscillation frequency is determined as
fo = fs + ∆ f (2.156)
The oscillation amplitude could be obtained from the following
mGx
Gm (x) = 1
(2.157)
1 − mα
where
C1 +C2
m= (2.158)
C1
and the crystal conductance is
R
Gx = 2 (2.159)
R + X2
With the above-mentioned procedure, given the crystal parameters (rs , Q0 , and fs ), one
can easily determine the oscillation frequency and the oscillation amplitude. The same
procedure can be used for a Hartley-like crystal oscillator.
2.13 Analysis of Large-signal Loop Gain of an Oscillator 81
Equation 2.160 describes the amplitude ratio as well as the phase difference. A more
general form of Equation 2.160 for the kth harmonic will be
It is possible to calculate the ratio of each harmonic to the main harmonic by Equa-
tion 2.161. In Equation 2.161, the load is impedance and is the one which is seen at
the collector of the transistor. Given the fact that harmonic currents are smaller than
the fundamental current and Q is large, it is obvious from Equation 2.161 that the
harmonic voltages are quite smaller than the fundamental voltage.
In order to extract the oscillator signal, we should not load the tank circuit di-
rectly because its quality factor will be degraded. It is possible to use either an
impedance transformer or extract the output from the emitter (where there is a low
output impedance). This point is illustrated in Figure 2.66.
As it is stated earlier, the output impedance of Figure 2.66 will decrease by p2 and
could reach to 50 Ω. An appropriate value of p (p < 1) will not degrade the quality
factor of tank circuit as much.
VCC VCC
1:P
RL rLoad RL LL
Q C1 Q C1
C∞
rLoad
Is C2 Is C2
(a) (b)
Figure 2.66: Extracting the output signal of an oscillator without degrading its
quality factor.
82 Chapter 2. Oscillators
where I0 (x) is the zeroth-order modified Bessel function of the first kind, IEQ is the
operating point DC current of the emitter (in the absence of the large-signal), and Vλ is
defined as
! !
αIEQ ln(I0 (x)) 2I1 (x) ln(I0 (x)) 2I1 (x)
Gm = 1+ Vλ
= gmQ 1 + Vλ
(2.164)
Vt xI0 (x) xI0 (x)
Vt Vt
As we increase RE , the voltage drop across it for a constant current will increase (Vλ
increases and the coefficient approaches unity), and as a result, it acts approximately
as a current source. The evolution of the transconductance of a bipolar transistor stage
biased with an emitter resistor instead of a current source is depicted in Figure 2.68.
VCC
CL RL LL
RB Vout
Q
Vss
RE
Vs
1
Vλ/Vt=20
0.8
Gm(x)/gmQ Vλ/Vt=40
Vλ/Vt=60
0.6 Vλ/Vt=80
Vλ/Vt=100
0.4
0.2
0 x
0 2 4 6 8 10 12 14 16 18 20
ID
VGS
VTH
I
Gm1= ~D1
VGS
Figure 2.69: Typical I−V characteristic of MOS transistor (ṼGS is the gate–
source’s AC voltage phasor).
84 Chapter 2. Oscillators
VDD
VGG CL RL LL
RFC
Vout
M
+
CG VGS0 -
V1cos(ωt)
I0 CS
Figure 2.70: Constant current MOS stage tuned amplifier for computation of
the large-signal transconductance.
where I0 is the current source’s bias current. Now, the large-signal transconductance is
defined as
I1 2 (VGS0 −VTH )
Gm = = I0 h (2.171)
V12
i
V1 (V −V ) + 2
GS0 TH 2
2.15 MOS Stage Large-Signal Transconductance 85
1
0.9
0.8
Gm(x)/gm
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0
0.5
1.5
2.5
3.5
4.5
5
x=V1/(VGS0-VTH)
Gm I0
= h (2.172)
V2
i
gm k (VGS0 −VTH )2 + 21
Or in another form
Gm I0
= (2.173)
V12
h i
gm 2
k(VGS0 −VTH ) 1 +
2(VGS0 −VTH )2
Given the fact that at the operating point, one can write
Gm 1
=h (2.175)
V12
i
gm 1+
2(VGS0 −VTH )2
V1
Let x = VGS0 −VTH ,
Gm (x) 1
= 2 (2.176)
gm 1 + x2
v 2 v
I2 = k VGS0 −VTH − for < |VGS0 −VTH | (2.178)
2 2
Then
2
I1 VGS0 −VTH + 2v
= 2 (2.179)
I2 VGS0 −VTH − 2v
Given
I1 + I2 = I0 (2.180)
VDD
LL RL CL CL RL LL
I1 I2
Vout
+
M1 M2
+
v
I0
iD2/I0 1
iD1/I0
0.9
0.8
0.7
0.6
I/I0
0.5
0.4
0.3
0.2
0.1
-2.5
-2
-1.5
-1
-0.5
0.5
1.5
2.5
V/(VGS0-VTH)
Figure 2.73: Variations of the differential pair drain currents as a function of
the normalized differential voltage in a MOS differential pair.
v
I0 VGS0 −VTH
I2 = 1− (2.184)
2 v2
1+ 2
4(VGS0 −VTH )
2, then I1 = I0 and I2 = 0, and if VGS0 v−VTH < −2, then I1 = 0 and I2 = I0 . As such, a
nonlinear transfer characteristic has been specified for a MOS differential pair for the
whole span of possible input voltages.
The differential pair small-signal transconductance becomes
I0
gmd = (2.186)
VGS0 −VTH
88 Chapter 2. Oscillators
V1
x= (2.188)
VGS0 −VTH
1 x cos θ
Z π
bn (x) = 2 cos nθ dθ (2.189)
π −π 1 + x4 cos2 θ
Care should be taken that these computations are valid for x ≤ 2. Note that given
the fact that the differential MOS pair transfer characteristic has an odd symmetry,
bn (x) functions would be zero for even values of n. The fundamental harmonic current
becomes
I1 = I0 b1 (x) (2.190)
I1 I0 b1 (x) b1 (x)
Gmd (x) = = = gmd (2.191)
V1 (VGS0 −VTH ) x x
1
0.9
Gmd(x)/gmd
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0
0.5
1.5
2.5
3.5
4.5
x=V1/(VGS0-VTH)
Figure 2.74: Normalized transconductance variations of a MOS differential
tuned amplifier stage as a function of the normalized input voltage.
2.17 An Oscillator With a Hypothetical Model 89
or
For the case where the large input signal does not satisfy the condition x ≤ 2, the
transistor pair drains would switch between zero and I0 . As such, the first harmonic
currents would have a value as
4
I1 = I0 (2.193)
π
The large-signal transconductance becomes
I0
I1 4 I0 4 VGS0 −VTH 4
Gmd (x) = ≈ = V1
= gmd (2.194)
V1 π V1 π VGS0 −VTH
πx
or
Gmd (x) 4
≈ (2.195)
gmd πx
√
This is valid for x ≥ 2 2.
As such, the overall normalized differential MOS stage transconductance is de-
picted in Figure 2.74.
This large-signal transconductance can be employed in the differential MOS
oscillator circuit design/analysis for computation of the amplitude of oscillation.
+\SRWKHWLFDO 9RXW P
. HOHPHQW 9L
,=
.
& / 5/
9=
9RXW
5LQN 9L
,V &
9ELDV
B2 2
IZ = B0 + B1Vi cos(ω0t) + V (1 + cos(2ω0t)) + B3Vi3 cos3 (ω0t) (2.197)
2 i
We are looking for a gain at the fundamental frequency in Equation 2.197. Thus,
one may obtain the first harmonic large-signal transconductance as
B1Vi + 34 B3Vi 3 3
Gm1 = = B1 + B3Vi 2 (2.198)
Vi 4
C1 +C2
Here, m = C1 . Or the oscillation condition becomes explicitly as
mRL Rin 3
B1 + B3Vi 2 = 1 (2.200)
RL + m2 Rin 4
and
1 C1C2 ω0
− =0 (2.201)
Lω0 C1 +C2
mGm
ALs ( jω) = j
= 1∠0 (2.202)
jCω − Lω + R1 + m2Yin
where Yin is the input admittance of the differential pair. The above equation describes
the closed-loop gain of the oscillator. Considering the fact that the input admittance of
2.19 Voltage-Controlled Oscillators 91
VDD ID
1:m
C R
VBias
id Vid
+ M2 M1
vid - Gm
Yin Is
Vid
the MOS differential pair is mainly capacitive, at the resonant frequency, regarding the
closed loop gain, one can write
1
Cω0 − + m2Yin ( jω0 ) = 0 (2.203)
Lω0
ALs ( jω0 ) ≈ mGm R = 1 (2.204)
5RWRU
SODQHV
6KDIW
6WDWRU
SODQHV
The capacitor is biased by a negative voltage where the reverse DC current of the
diode is negligible. The negative voltage should be less than the breakdown voltage of
the diode junction. Moreover, it is possible to switch between capacitors to change the
frequency coarsely. Figure 2.80 shows the implementation of a variable capacitance in
an oscillator.
At low frequency, the bias current of the variable capacitors passes thorough the
inductor (note that the inductor is short-circuit at low frequencies). The oscillation
frequency of the circuit is determined by the resonance of the total capacitance of
VDD
L1 L2
M3 M4
M2 M1
C2 C3
IDC
R
VC
C1
C1 and C2 plus the capacitances of the varactor with the inductor, L. Furthermore, it
is possible to modulate the frequency of oscillation by a time-varying voltage at the
cathode of the varactors. The bias resistance of the varactors (Rbias ) is assumed to
be sufficiently large in order to avoid the loading of the tuned circuit or else an RFC
should be added in the bias circuit. At high frequencies, the varactor capacitances are
in series.
As Figure 2.79 suggests, the variable capacitor is nonlinear which may result in
signal distortion (generation of RF harmonics). In Figure 2.80, the oscillation voltage
is divided between the two varactors, resulting in better linearity. Indeed, a varactor
pair allows for double AC voltage swing, and thus, a lower distortion can be achieved
by a varactor pair at the output (with respect to a single varactor oscillator). In fact, the
bias voltage across these varactors changes their values and these changes will result
in resonant frequency variations. Three different kinds of VCOs with their varactor
implementation are shown in Figure 2.81.
In Figure 2.81(a), the tuning voltage can select any value greater than zero and vari-
able capacitance will experience the substrate noise. Tuning voltage in Figure 2.81(b)
can have only values smaller than supply voltage for staying in reverse bias. It also
experiences the supply voltage noise and its ripples. Finally, Figure 2.81(c) will have
the same tuning voltage as Figure 2.81(b). In this case, from the supply point of view,
the two varactors are in parallel; however, from the RF point of view, those are in series,
therefore, the supply noise voltage is canceled out in the RF circuit. Furthermore, the
nonlinear behavior of the VCO is ameliorated in this case.
&
2SHUDWLRQ
9'
UHJLRQ
4 &
/ 5 5ELDV
9&&9&RQWURO
5HYHUVH
& ELDV
,V
9&& 9&&
Q C1 C∞ Q C1 Q C1
VBias
Vbias Vbias Vbias
Is C2 VC1 Is C2 Is C2
(a) (b) ( c)
Example 2.9 Consider the given differential MOS pair in the circuit of
Figure 2.82.
I1 I2
V1 M1 M2 V2
ISS
(b) Now, suppose that the input is a differential signal with Vi = Vm cos (ωt). Find
the expansion of Equation 2.206, and then find the large-signal transconductance.
(Large-signal transconductance is the ratio of the fundamental harmonic current to
the input voltage amplitude.)
(c) Now, suppose that with the circuit of part (a), we have implemented the oscillator
of Figure 2.83.
With the assumption of ideal transformer and with the derived equation for the
large-signal transconductance, find the loop gain and then the oscillation criteria.
Finally, with R = 3 kΩ, k = 1 m VA2 , and ISS = 1 mA, find the coupling coefficient n
for the oscillation amplitude of 600 mV.
96 Chapter 2. Oscillators
VDD
1:2n
Vb
R L C
M1 M2
ISS
(d) Suppose that this oscillator is designed to oscillate at 1 GHz with the values
of L = 5 nH, and C = 5 pF. To change the oscillator, to the VCO of Figure 2.84,
we need to add the varactors in parallel to the capacitors. The characteristic of the
varactor is shown in Figure 2.85.
VDD
1:2n
Cv
Vcont Vb
R L C
Cv
M1 M2
ISS
0.5pF
Vvar
-VDD
2
I1,2 = k VGS1,2 −VTH (2.207b)
As Equation 2.216 suggests, when the oscillation amplitude increases, the large-
signal transconductance decreases which results in stable oscillation.
(c) As we stated earlier, for the oscillation condition, the resonant circuit impedance
will become real at the oscillation frequency. Now, bearing this in mind, we write
the loop gain. The operation of the oscillator is as follows, the differential pair
converts the input voltage to the output current at the opposite drain, and then this
current flows through the resonant circuit and generates the output voltage. Finally,
the transformer returns a part of the output voltage to the input with the same phase
(positive feedback). Therefore, one may write the loop gain (assuming that the
input of the MOS stage does’nt load the output transformer), HL , as
1 1
fmax = r = 982.3 MHz (2.218)
2π CV,min
L C+ 2
2.19 Voltage-Controlled Oscillators 99
1 1
fmin = r = 918.8 MHz (2.219)
2π CV,max
L C+ 2
(e) Using Equation 2.215, we can compute the third harmonic from Equation 2.215.
Then, the ratio of the third harmonic to the first one becomes
|H3 | kVm 2 3
= 2
≈ 9.5 × 10−5 or − 80.4 dBc (2.220)
|H1 | 16ISS − 3 kVm 8Q
Example 2.10 Consider the oscillator circuit depicted in Figure 2.86 which is a
VCO.
$PSPRGHO
9RXW
Y 5LQ L2 52 & &Y
5)& 5%
9LQ 9E
/
&Y &
&
CV
145pF
120pF
105pF
Vvar
-3 -2 -1
Figure 2.87: Varactor characteristics.
100 Chapter 2. Oscillators
Note that
3 cos θ + cos 3θ 10 cos θ + 5 cos 3θ + cos 5θ
cos3 θ = , cos5 θ = (2.224)
4 16
(a) Investigate the effect of finite resistance of RFC if its equivalent parallel resistor
is 2.5 kΩ.
(b) Find the amplitude and the frequency of oscillations for Vb = 1 V.
(c) Find the range of oscillation frequency for 1 < Vb < 3.
(d) Find the amplitude of the fifth harmonic in case (a).
Solution:
(a) The RFC resistance is added in parallel to the output. To calculate its effect, we
should first determine the total output resistance at the operating frequency. The
effective loading resistance of the RFC which would appear at the output of the
oscillator would be
0
RPRFC = n2 RPRFC = 10 kΩ (2.225)
where n = 2 because of the existence of the double varactors. The effect of the
RFC parallel resistance would be to reduce the output resistance and the overall
gain as described in part (b).
(b) The total capacitance at the output (at Vb = 1 V) is computed as
C1C2 CV
CT = + = 207.5 pF (2.226)
C1 +C2 2
1
f= √ = 110.487 MHz (2.227)
2π LCT
where
0
RT = Ro k RPL k RPRFC = 314.3 (2.231)
By expanding Equation 2.232 and considering only the main harmonic component
of the current, we obtain
3c 2 5e 4
io1 = V1 cos (ω0t) a + V1 + V1 (2.234)
4 8
3c 2 5e 4
Gm 1 = a + V1 + V1 = 0.0318 (2.235)
4 8
Resolving Equation 2.235, one obtains two possible solutions:
V1 = 0.197 V (2.236a)
or
V1 = 2.63 V (2.236b)
∂ Gm
Only one of these solutions is acceptable and that is the one for which ∂V 1 < 0.
1
Actually, the loop gain should decrease with the amplitude at a stable point (why?),
and here only the first solution has such a characteristic.
(c) For the frequency range of the VCO, we can compute the upper bound
and the lower bound frequency of the oscillation from Equation 2.227 for CVmin =
105 pF and CVmax = 145 pF, respectively. Therefore, the frequency range will be
from 116.230 MHz to 110.487 MHz .
102 Chapter 2. Oscillators
1
V5 = 0.197 × = 2.7 µV (2.240)
73110
Equating Equations 2.241 and 2.242, one obtains a relation between the base–emitter
voltage and the emitter current.
2.20 Special Topic: Nonlinear Device Fed by Sinusoidal Large-Signal 103
kT IEQ kT IE1
vBE = ln + ln 1 + cos (ωt) (2.243)
q IES q IEQ
Or in another form
kT
vBE = VBEQ + ln [1 + y cos (ωt)] (2.244)
q
where
IE1
y= (2.245)
IEQ
As such, the large-signal input resistance seen from the emitter can be obtained
p
VBE1 VBE1 IEQ kT 2 1 − 1 − y2
Rin (y) = = = × (2.247)
IE1 IEQ IE1 qIEQ y2
rin y2
= (2.248)
Rin (y) 2 1 − 1 − y2
p
Vout
Q
C1
IE1
L RL
RE IEQ C2
-VEE VCC
Figure 2.88: A crystal Colpitts oscillator (Butler oscillator) where the input
current at the emitter is approximately sinusoidal.
104 Chapter 2. Oscillators
where rin is the small signal input resistance seen through the emitter:
kT Vt
rin = = (2.249)
qIEQ IEQ
The variations of the inverse normalized emitter resistance as a function of the normal-
ized input current are depicted in Figure 2.89. Note that 0 < y < 1 in essence.
1
0.95
0.9
0.85
rin/Rin
0.8
0.75
0.7
0.65
0.6
0.55
0.5 y
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
Vout
Si C1=300pF
α=0.98
L RL=5kΩ
RE=8.4kΩ C2=15nF
10MHz
-VEE VCC
In this circuit, if the resonant frequency of the tank circuit and the crystal
resonant frequency are not exactly the same, a slight change in frequency and the
amplitude will occur. Find the amplitude and the frequency of the oscillations for
two cases: (i) determine the value of inductance L for the oscillations at 10 MHz,
(ii) if the resonance frequency of the tank circuit is reduced by an amount of 10 kHz,
for example, due to temperature or process variations, determine the new frequency
2.20 Special Topic: Nonlinear Device Fed by Sinusoidal Large-Signal 105
Solution:
First, we assume that both resonant frequencies are the same, and therefore we
attain
C1 1
n= = (2.250)
C1 +C2 51
C1C2
Ceq = = 294 pF (2.251)
C1 +C2
1
L= = 861 nH (2.253)
Ceq ωs 2
and
VEE −VBEQ
IEQ = = 0.512 mA (2.254)
REE
and
kT
rin = = 50.7 Ω (2.255)
qIEQ
In Figure 2.91, the current loop gain at the resonant frequency can be written as
αIE1 RL n
= IE1 (2.256)
rs + Rin
Note that for n 1, the secondary of the transformer in Figure 2.91 doesn’t load
the primary. The oscillation condition is
αRL n
=1 (2.257)
rs + Rin
Therefore
Using Figure 2.89, one obtains y = 0.75. Therefore, for the oscillation amplitude,
we attain
Now, suppose that due to process variation, the resonant frequency of the tank is
10 kHz lower than the series resonance frequency of the crystal, i.e., fo = fcrystal −10
kHz. The tank circuit introduces a phase-change which must be compensated by
the crystal. Since the rate of change of the reactance of the crystal is extremely
high near the resonant frequency, a slight change in the frequency will compensate
the aforementioned phase shift. To calculate the oscillation frequency alongside
the oscillation amplitude, consider Figure 2.91.
rs Ls Cs IE1
1:n
αIE1 L RL C Rin
Figure 2.91: Crystal Colpitts oscillator equivalent circuit seen through the
emitter.
The oscillation criteria mandate that the overall current loop gain to be unity
with zero phase, therefore we obtain (neglecting the loading of the secondary
impedance on the primary of the transformer, given the fact that n 1)
αIE1 RL n 1
× = IE1 (2.262)
ω
1 + jQt ω0 − ω ω0 rs + R in + jX
Or
αRL n 1
× = 1∠0 (2.263)
ω
1 + jQt ωω0 − ω0 rs + Rin + jX
2.20 Special Topic: Nonlinear Device Fed by Sinusoidal Large-Signal 107
where in Equation 2.262, X is the crystal reactance and Rin is the emitter input
resistance. Therefore, from Equation 2.263, we have
ω − ωs ω − ωs
X ≈ 2QL (rs + Rin ) = 2Q0 rs (2.264)
ωs ωs
where QL is the loaded quality factor and Q0 is the unloaded quality factor of the
crystal. The magnitude of Equation 2.263 can be written as
(αRL n)2 1
2 × =1 (2.265)
(rs + Rin )2 + X 2
1 + Qt ωω0 − ωω0
αRL n
rs + Rin = 2 (2.268)
1 + Qt ωω0 − ωω0
αRL n
Rin = 2 − rs ≈ 57.9 Ω (2.269)
ωs ω0
1 + Qt ω0 − ωs
fs
∆f = X ≈ −63 Hz (2.271)
2Q0 rs
Now for computing the new oscillation amplitude, we should consider the new
value for Rin . The inverse normalized large-signal resistance becomes
rin
= 0.875 (2.273)
Rin (y)
Using Figure 2.89, one obtains y = 0.66. The output tuned circuit voltage can be
calculated as
RL
|vt | = αIE1 |ZL | = αyIEQ r 2 = 1.69 V (2.274)
1 + Q2t ω
ω0 − ω0
ω
• Electrical Specifications
Min. 2315
Frequency (MHz) Max. 2536
Power output (dBm) Typ. +6
1 −75
Typical phase noise (dBc/Hz) 10 −105
SSB at offset frequencies, kHz 100 −128
1000 −148
Min. 0.5
Tuning voltage range (V) Max. 5
2.21 Datasheet of a Voltage-Controlled Oscillator 109
• Performance Data
• Curves
2XWSXW3RZHUG%P
)UHTXHQF\0+]
+DUPRQLFVG%F
Figure 2.92: Oscillation frequency, output power, and harmonic levels of the
ZX95 VCO as a function of the tuning voltage.
110 Chapter 2. Oscillators
• Curves
2XWSXW3RZHUG%P
)UHTXHQF\0+]
+DUPRQLFVG%F
)
)
)
&
&
&
7XQLQJYROWDJH9 7XQLQJYROWDJH9 7XQLQJYROWDJH9
Figure 2.93: Oscillation frequency, output power, and harmonic levels of the
POS-100 VCO as a function of the tuning voltage.
2.22 Conclusion
In this chapter, we have studied the basic operation of sinusoidal oscillators. Oscillators
generally operate by means of amplification of circuit noise in a relatively high-gain
frequency selective closed-loop circuit. Here, the noise as an initial signal contributes
to the build up of the oscillator sinusoidal signal and the loop gain of the oscillator
is consequently compressed (reduced) by the generated large signal. For a stable
oscillation, it is necessary to satisfy Barkhausen’s criteria. That is to say, to achieve a
unity closed-loop gain with 2π or zero phase. In general, an active element in addition
to a frequency selective (resonating) circuit is needed in an oscillator. It is noteworthy
that the oscillators generally operate in large-signal regime. So it is important to have
a nonlinear model for the device in order to compute adequately the amplitude and the
frequency of the oscillation.
In this chapter, different oscillator topologies including CE, CB, CC, (or CS, CG,
and CD for MOS transistors) as well as Colpitts-like or Hartely-like oscillators were
studied. The study of oscillator circuit is essentially divided into two parts: In the
first part, the resonant dividing circuits were studied where RLC resonant circuits are
used with either capacitive or inductive dividers. In the second part, the nonlinear
behavior of the active elements used in the oscillator circuits should be studied. Here
we presented large-signal models for the bipolar transistors, differential bipolar pairs,
MOS transistors, and the MOS differential pairs where large-signal transconductances
were computed for either of the active elements. Using variable capacitors or varactors
in the oscillator circuits permits the frequency tuning of them. As such, voltage-
controlled oscillators (VCO) were presented. VCOs are one of the main building
blocks of the phase-locked loops which will be discussed in the next chapter. The
main figures of merits of an oscillator are its frequency stability, its spectral purity, low
harmonic level, and its low phase noise.
2.24 Problems
Problem 2.1 Consider the resonant circuit depicted in Figure 2.94 which is normally
used in Clapp oscillators.
1. We know that the oscillation will occur where the impedance of resonant circuit
is pure real which corresponds to zero phase shift. In the given resonant circuit,
find at which frequency the impedance will be pure real?
L
Zin CP R
CS
2. With the results of part 1, find the oscillation frequency of Figure 2.95 for
L1 = 50 nH and C0 = C1 = C2 = 3 nF.
VDD
RFC
VGG
L1 C1
C0 I
C2
Problem 2.2 Consider the Colpitts oscillator depicted in Figure 2.96 with the given
values of parasitic capacitances, namely, base–collector Cµ = 15 pF, collector–substrate
CCS = 15 pF, base–emitter Cπ = 30 pF, IC = 3 mA, and CB is RF short.
VCC= 10V
R1 L1 RL
7.8kΩ 20nH 500Ω
Vout
C1
230pF
R2
CB
2.2kΩ I C2
3mA 200pF
VDD
RL
R1 L1
830Ω
Vout
C1
Cg R2 I=
C2=C1
2mA
Problem 2.4 For the given Colpitts oscillator depicted in Figure 2.98 which is
supposed to oscillate at 100 MHz ? Consider the transistor’s Early voltage is 40 V and
β = 100.
3v
Q=? Cµ
L=? 0.2pF 5mA
Q=100 Cπ
C2=20pF 2pF
Q=100
RE=?
C1=20pF
5v
VA=40v
L1=300nH
6kΩ 0.1pF β=100
Q=20
Vout
C1=?
1pF Q=100
Problem 2.6 Figure 2.100 depicts the block diagram of a hypothetical oscillator.
5HVRQDQW
/LQHDU
$03 FLUFXLW
/LPLWHU
Each block in Figure 2.100 can be modeled by an ideal element (for the amplifier,
you may put a voltage-dependent current source with gm transconductance and finite
output resistance R, and for the resonant circuit, an LC resonator with infinite quality
factor). Suppose that the limiter characteristics follow Vo = tanh (bVin ) where Vin and
Vout are the input and the output voltages of the limiter, respectively. Moreover, assume
we have a gm = 4 mf, R = 500 Ω,C = 5 pF, L = 5 nH, and |b| = 10 V−1 .
1. Draw an equivalent circuit diagram for the oscillator, and determine the oscilla-
tion frequency.
2. Determine the effective gain of the limiter.
3. Determine the sign of the parameter b in the characteristic of the limiter for
positive feedback.
Problem 2.7 Common topologies of the MOS oscillators are shown in Figure 2.101.
All transistors have the transconductance and their parasitic capacitances are
Cgs = 231 fF,Cgd = 94 fF,Csb = 24 fF, and Cdb = 19 fF. Moreover, for other parameters,
we have L = 1.5 nH,C1 = 20 pF, and C2 = 5 pF. With the given values, find the
oscillation frequency and compare them in the three topologies.
2.24 Problems 117
Problem 2.8 In the circuit depicted in Figure 2.102, assume that the input large
signal of the stage is Vi = Vm cos (ω0t), and the I − V characteristics of the active
2
device follow I = 12 K Vgs −Vth for Vgs ≥ Vth , and I = 0 for Vgs ≤ Vth . Assume
QL = 50, the output circuit is tuned to ω0 and Vb = Vth . Find the conduction angle in
the output current and then find the first to the fifth output current harmonics and the
first to the fifth output voltage harmonics.
VDD
L C
Q=100
M1
Vb
+
Vi CB
Problem 2.9 The crystal oscillator depicted in Figure 2.103 is named after its de-
signer as Driscoll oscillator. In this oscillator contrary to other types of oscillators, at
the oscillation condition, transistor Q1 will not be driven to the nonlinear regime and
the diodes D1 and D2 will be driven to the nonlinear region and as such will limit the
signal level. Using the exponential I −V characteristics of diodes and using the Bessel
function expansion, find the loop gain, and find the amplitude of the oscillation (an
important feature of the circuit in Figure 2.103 is the separation of the resonant circuit
from the limiter which results in better phase noise of the oscillator). Assume that Vb1
and Vb2 are adequate positive voltages to maintain Q1 and Q2 in their active region.
Furthermore, the phase shifter block has a voltage gain of unity and it doesn’t load the
output tuned circuit.
VCC
180º phase shift at L2 C2 RL D1 D2
oscillation frequency
L1 C∞
Vout
C1 C1
Q2
Vb2
Rb
Q1
C∞
Vb1
Tuner of the crystal RE
capacitance
Problem 2.10 Assuming a square law characteristics for the MOS transistors as
in Equation 2.275, one can derive the drain current of the MOS differential stage as
Equation 2.276. Using the polynomial expansion of Equation 2.276, find the large-
signal transconductance of the stage and the large-signal loop gain to deduce the
amplitude of oscillation.
1 W 2
ID = µnCox Vgs −Vth (2.275)
2 L
s
ISS µnCox W 4ISS
ID1 = −Vg2 −Vg2 2 (2.276)
2 4 L µnCox WL
VDD
C1
R L
C2
RFC
M2 M1
+
Vg2 - ISS
-VSS
Problem 2.11 Consider the Colpitts oscillator depicted in Figure 2.105 and assume
that Vcontrol = 6 V, bipolar transistor’s β = 100, L = 2 µH,C1 = 55 pF, and C2 = 550 pF.
Moreover, Figure 2.106 depicts the variable capacitance characteristics. Assume that
the MOS transistor is off.
VCC=3V
RL C2 CV1 Rb
Vs M1 L Vcontrol
C1 CV2
R1 C∞
Vout
356fF
1pF
C∞ R2
IE=2.5mA RE=500Ω
CV1,2
3.25pF
2pF
0.85pF
-7 -3 -1 Vvar
9&& Y
5/
,GHDO 9RXW
(QYHORSH
GHWHFWLRQ
P9'&
442pF
1μH
442pF
Vin
Rin= Rout=
250Ω +
Vo 1000Ω
Problem 2.13 In an oscillator depicted in Figure 2.109 where the nonlinear transfer
characteristics of the device are given by Equation 2.278, find the oscillation frequency
as well as the amplitude of the fundamental and the second harmonic.
L&
UR *P9EH S)
ȍ
Q+
9EH UH
UH§*P ȍ 9RXW
S)
9ELDV ,ELDV
9ELDV
v1 = vbe (2.277)
1 v1 v1 2 v1 3
iC = + + − A (2.278)
100 50 300 400
Problem 2.14 Design problem. In the reference Driscoll oscillator depicted in
Figure 2.110,
1. How the values of C1 , C2 , L4 , and L5 are determined, in such a way that the
circuit oscillates at the third series resonance of the crystal ( fS ). Describe the
corresponding relations (write the oscillation condition).
2.24 Problems 121
2. The oscillation amplitude at the emitter of Q1 is much smaller than the thermal
voltage VT , and at the collector of Q2 , it is larger than a few VT ’s, and the third
overtone of the crystal is the dominant impedance. The amplitude of oscillations
at the collector of Q2 is determined by the Schottky diodes impedances. The
VD
I −V characteristic of diodes follows ID = IS e Vt . Find the harmonic content of
the current by the Bessel function expansion of the output characteristics and
determine an expression for the loading impedance of the diodes.
VCC
Vo 50Ω
L4
1 C∞
Rb3 20
L5
C2
Q2
C∞
Rb2 C∞
C∞ R R C∞
Q1
-VC +VC
RFC
Rb1 C1
REb
Problem 2.15 For the oscillator depicted in Figure 2.111, using the large-signal
model of the transistor, draw the equivalent circuit and write the complex relation
describing Barkhausen’s oscillation criteria. Assume that the crystal’s admittance is
represented by a complex value Yx . Note: do not use the equivalent transformer model
for the capacitive divider in this case.
VCC
R1
Si
α=0.99
C1
Yx R2
RE
C2
Problem 2.16 Writing the oscillation condition, find the frequency and the amplitude
of oscillations in the circuit of Figure 2.112 at the collector of the transistor Q1 .
Furthermore, find the third harmonic amplitude at the same node.
5V M12
L1 L2
Vout
Si
Q1 Q2
α=0.99
R+jX
+VCC
+ ro
Ri µVi
Vi +
C1 C2
-VCC
Problem 2.18 In the Pierce crystal oscillator depicted in Figure 2.114, the crystal
is inductive and will resonate with the input and the output capacitors resulting in a
sinusoidal oscillation. Assume the nonlinear element has the given I −V characteristics.
First find the large-signal effective Gm as a function of AC voltage amplitude and draw
it to the scale. Then, write the complex oscillation condition in Figure 2.114 as a
function of the circuit parameters. What is the required Gm for an oscillation amplitude
of 1 V?
2.24 Problems 123
5M;
*PPV
5 & &L L I9 & 5
9
L Į9ܵ9ı9_9_ 9
Į P$9
99 ܵ
P$9
ı P$9
4 Mixers . . . . . . . . . . . . . . . . . . . . . . . . 169
4.1 Mixer Concept
4.2 Third Order Intermodulation Concept in a Nonlinear
Amplifier
4.3 Basic Concept of Third-Order IM in a Basic Mixer
4.4 Bipolar Transistor Active Mixer
4.5 Mixer types Based on Switching Circuits
4.6 Matching in Mixers
4.7 Calculating IIP3 in Nonlinear Amplifier/Mixer
4.8 Linearization Methods in Mixers
4.9 Calculating Third-Order Input Intercept Point in
Cascaded Stages
4.10 Important Point in RF Circuit Simulation
4.11 Conclusion
4.12 References and Further Reading
4.13 Problems
5 Modulation/Demodulation of Amplitude/Phase
223
5.1 AM Modulation
5.2 AM Demodulation
5.3 Generating AM Signals
5.4 Double-Sideband and Single-Sideband Suppressed
Carrier Generation
5.5 Synchronous AM Detection
5.6 Gilbert Cell Applications
5.7 Modern Practical Modulations
5.8 Effect of Phase and Amplitude Mismatch on the Signal
Constellation
5.9 Conclusion
5.10 References and Further Reading
5.11 Problems
9&&
&YY
&YY S) S)
/
9' 9 9
4 S)
Q) S) &9W
&
S)
0LFURSKRQH 5 5 9'
S) ,6
9 9
&
9'W
9'
FRQWUROYROWDJH
Figure 3.1: A typical voltage-controlled oscillator (VCO) with voice input for
frequency modulation.
where Vm is the maximum value of the baseband voltage, VB is the varactor’s bias
voltage, and g (t) is a random normalized function (varying between +1 and −1)
proportional to the input information. The function g (t) may be a continuously valued
analog signal for FM or a two discrete level valued digital signal for FSK modulation.
The instantaneous frequency can be computed as
!
1 C(t)
1− 2 C1C2 +C (V )
1 C +C
1 2
v B
f= r ≈ r = f0 + ∆ f (t)
C1C2 C1C2
2π L C1 +C2 +Cv (VB ) +C (t) 2π L C1 +C2 +Cv (VB )
(3.3)
1
f0 = r (3.4)
2π L CC11+C
C2
2
+Cv (VB )
Here the carrier frequency and the frequency deviation are clearly described as a
function of circuit parameter values. Now with the definition of frequency modulation,
3.1 Frequency Modulation 129
We can obtain the instantaneous frequency of Equation 3.6 by the derivation of the
argument of the sinusoidal signal:
dΘ
ω (t) = (3.7)
dt
where in Equation 3.7, Θ is called the total phase. Moreover, we know that
where in Equation 3.8 ω0 is the carrier radian frequency and ∆ωm (t) is a function of
the baseband or the radian frequency deviation. Comparing Equations 3.7 and 3.8, we
reach to
Z Z
Θ= (ω0 + ∆ωm (t))dt = ω0t + ∆ωm (t)dt (3.9)
Thus far, we have calculated the signal phase for Equation 3.6 and also introduced
the frequency modulation. As Equation 3.6 suggests, we have defined the frequency-
modulated signal by a constant amplitude sinusoid. The information is embedded in
∆ωm (t) which changes the VCO frequency proportionally. It should be noted that
the bigger the amplitude of the input baseband signal the more will be the frequency
deviation. Thus, in many applications, we employ a limiter in the baseband circuit
to limit the bandwidth occupancy. Therefore, the information is merely in a specific
bandwidth. As an example, we can inspect the specifications of the commercial FM
radio. The standard obligates the bandwidth of this radio to be limited to 200 kHz. The
Carson bandwidth equation predicts that
BW FM ∼
= 2 ( fdev + fBB ) (3.10)
where in Equation 3.10, fdev is the frequency deviation and fBB is the maximum
baseband signal frequency. As an example, for a frequency deviation of 75 kHz, and a
maximum baseband frequency of 20 kHz, the Carson bandwidth becomes
BW FM ∼
= 2 (75 + 20) = 190 kHz (3.11)
which is within the specified bandwidth of 200 kHz. As another example, the time
variation of an FM signal is illustrated in Figure 3.2. Here for the sake of illustration,
the carrier frequency is chosen as 500 kHz and the frequency deviation is chosen as
75 kHz.
As depicted in Figure 3.2, when the amplitude of the baseband signal is high, the
frequency increases and when this amplitude is low, the frequency decreases.
130 Chapter 3. PLL, FM Modulation, and FM Demodulation
%DVHEDQG
VLJQDO
N+]
N+] N+]
)UHTXHQF\
PRGXODWHG
VLJQDO
Figure 3.2: Typical baseband and the corresponding FM signal variations with
time.
Z t
dvFM
= A (ω0 + ∆ωm (t)) cos ω0t + ∆ωm (t) dt (3.12)
dt 0
The above signal can be detected using an envelope detector. As such, it is possible
to detect an FM signal through a differentiator followed by an AM detector. We now
continue with the other concepts for frequency demodulation.
1
V4
V1 0
1 1
V2 0
∆ɸ -π π ∆ɸ
1
2π V1 V3
V3 0 LPF V4
V2
Figure 3.3: XOR (phase detection) characteristics driven by two signals with a
common frequency and a fixed phase shift.
I0 qv
1
IC1,C2 = 1 ± tanh (3.13)
2 2 kT
VCC VDD
RL - vout + RL RL - vout + RL
IC3,5 IC4,6 ID3,5 ID4,6
CL CL
Q3 Q4 Q5 Q6 M3 M4 M5 M6
v2 v2
I0 I0
(a) (b)
Figure 3.4: Gilbert cell (analog multiplier), (a) by bipolar transistor pairs and
(b) by MOS transistor pairs, used as a phase detector.
132 Chapter 3. PLL, FM Modulation, and FM Demodulation
For large signal inputs, that is, VV1t 1 and VV2t 1, the hyperbolic tangent of sinusoidal
signals turn into square-wave signals of the same frequency and phase. That is
Vout = I0 RL S (ω0t) S (ω0t + φ ) (3.16)
1
∆ɸ =0 0
-1
Vout,LP 1
∆ɸ =π/4 0
RLI0 -1
1
-π π ∆ɸ =π/2
∆ɸ 0
-1
1
-RLI0 ∆ɸ =3π/4 0
-1
1
∆ɸ =π 0
-1
(a) (b)
Figure 3.5: Gilbert cell function as a phase detector. (a) The phase detector
output characteristics. (b) Signal waveforms for phase detector operation.
3.2 Frequency Demodulation 133
where S (ω0t) is a bipolar square wave of a unity amplitude. The Fourier expansion of
the square waves gives
4 1 1
Vout = I0 RL cos (ω0t) − cos (3ω0t) + cos (5ω0t) − · · ·
π 3 5
4 1 1
× cos (ω0t + φ ) − cos (3ω0t + 3φ ) + cos (5ω0t + 5φ ) − · · · (3.17)
π 3 5
Here we assume that vGS1 = VGS01 + v1 /2 and vGS2 = VGS01 − v1 /2. The ratio of
the currents in the lower tree transistors, with the above assumption, becomes (with
v1
2 < VGS01 −VTH to remain in the square law region)
2 2
I1 VGS01 + v21 −VTH Veff + v21
= 2 = 2 (3.20)
I2 VGS01 − v21 −VTH Veff − v21
I1 + I2 = I0 (3.21)
and the current in the either drains of the lower tree transistors can be described as
v1 v1
I0 VGS01 −VTH I0 Veff
I1 = 1 + 2 = 1 + (3.22)
2
2 2
1 v1 1 v1
1+ 4 VGS01 −VTH 1+ 4 Veff
v1 v1
I0 VGS01 −VTH I0 Veff
I2 = 1 − 2 = 1 − (3.23)
2
2 2
v1 v1
1 + 41 VGS01 −VTH 1 + 41 Veff
134 Chapter 3. PLL, FM Modulation, and FM Demodulation
where, VGS01 is the DC bias voltage of the lower tree transistors and VGS02 is the
DC bias voltage of the upper tree transistors. Here again, the output voltage will be
proportional to the low-pass component of RL ∆I. Note that the MOS Gilbert cell
produces a differential current proportional to the analog multiplication of the input
voltages approximately. This function is again a saturating function of the input
voltages and tends approximately to ±1 once its argument (input voltage) is large. As
such, again the output current will exhibit a bipolar XOR (Exclusive-OR) function of
the two large inputs (an Exclusive-OR with ±1 logic levels). This way the low-pass
component of the output will be proportional to the phase difference of the inputs,
provided that the inputs are large signal, that is, larger than 2Veff . The same analytical
procedure as described in Equations 3.16 through 3.18 holds for MOS phase detector as
well and consequently, a triangular output characteristics is produced here again. The
phase detector characteristics, that is, the output voltage versus the phase difference of
the two input signals, are shown in Figure 3.5(b).
In general, the Gilbert cell will be a phase detector when both inputs are driven to
the large signal regime. If the lower tree is driven by a small signal and the upper tree
experiences hard switching, the circuit changes to a mixer operation (which we will
study in Chapter 4). However, if both trees are driven by small-signal inputs, we will
have an analog multiplier. Now we desire to design a phase detector with a Gilbert
cell. We know that the data lie in the phase or the frequency of the carrier.
5DGLRVLJQDO
0+] $PSOLILHU ([WHUQDOILOWHU /LPLWHU
%3) %3)
0+] 0+]
*LOEHUW
0+] /3) FHOO
9RLFH 9
VLJQDO
&6
9
)UHTXHQF\
GHPRGXODWRU /3 &3 53
detector. Here, a Gilbert cell can be used as the phase detector. Assuming a high input
impedance for both inputs of the phase detector, we can easily show that
V2 CS LP s2
= (3.25)
V1 s2 S
ω12
+ Qω 1
+1
where
1
ω1 = p (3.26a)
LP (CP +CS )
Q = RP (CP +CS ) ω1 (3.26b)
Here Q is the detector’s quality factor and ω1 is the center frequency of the detector.
In the frequency domain, one can write
V2 jQ CSC+C
S ω
jQ CSC+C
S
= P ω1 ≈ P
(3.27)
V1 1 + jQ ωω1 ω ω2
− 1 1 + j2Q ∆ω
ω
2 1
1
V2
V1
QCS/(CP+CS) High Q
Low Q
f
f1
V2
V1
180º
90º f1
0º f
ω = ω1 + ∆ω f (t) (3.30)
where f (t) is the normalized baseband signal, that is, −1 ≤ f (t) ≤ +1. The approxi-
mate formula for Equation 3.28 can be written as
ω1
− (ω1 +∆ω f (t))Q π 2Q
∆φ = tan−1 2 ≈ − (∆ω) f (t) (3.31)
2 ω1
ω1
1 − ω +∆ω f (t)
1
As it is seen in Equation 3.31, the phase difference between the two voltages V1 and
V2 is proportional to the instantaneous frequency deviation ∆ω f (t) (while the two
voltages are at quadrature at the resonant frequency). Or in other words, as Figure 3.7
suggests, the phase difference between two inputs of phase detector at f1 is 90◦ . Due
to the frequency deviation of the frequency-modulated signal, the output voltage varies
with frequency deviation and consequently with the slope of the phase characteristics
of the quadrature tank. Figure 3.8 depicts the phase characteristics of the quadrature
tank for different quality factors.
As it is obvious from Figure 3.8, for a higher quality factor, we will attain a higher
sensitivity for a specific frequency deviation and the characteristic of Figure 3.8 will
be sharper. We can approximate the phase variations in Figure 3.8 near the center
frequency linearly as depicted in Figure 3.9.
3.2 Frequency Demodulation 137
V2
V1
High Q
180º
Low Q
90º f1
0º f
9
9 +LJK4
/RZ4
I
I
)UHTXHQF\
GHYLDWLRQ
This linear approximation is valid for a small frequency range. In our example, the
IF frequency is 10.7 MHz which is normally used in FM receivers. Moreover, as stated
earlier, the frequency deviation is ±75 KHz. The mentioned frequency deviation is
the maximum value; however, its instantaneous value depends on the input baseband
signal. We can write the phase difference in Figure 3.9 with Equation 3.31 as
π 2Q
∆φ = − (∆ω) f (t) (3.32)
2 ω1
where in Equation 3.32, ∆ω is the frequency deviation, f (t) is the voice signal, Q is
the quality factor, and ω1 is the center frequency. The quality factor is the following
Q = (CS +CP ) ω1 RP (3.33)
In this circuit, the frequency deviation is translated to ∆φ and the phase detector
translates the phase difference to a voltage proportional to the baseband. Now, let’s
analyze quantitatively the output from the multiplication occurring in the phase detector.
Consider the FM input applied to port1 of the phase detector as
Z t
V1 = A cos ω1t + ∆ω f (τ)dτ (3.34)
0
138 Chapter 3. PLL, FM Modulation, and FM Demodulation
It should be noted that higher quality factor results in more nonlinearity and distortion
in the detection process. For a moment if we assume the phase detector as a multiplier
and with the inputs of Equations 3.34 and 3.35, then the output would have a form like
QA2 CS
π 2Q
V1V2 = cos − (∆ω) f (t)
2 CS +CP 2 ω1
Z t
π 2Q
− cos 2ω1t + 2∆ω f (τ)dτ + − (∆ω) f (t) (3.36)
0 2 ω1
But considering the amplitude A is large enough to make a hard switching for the upper
tree of the Gilbert cell, and the amplitude of V2 that is QA CSC+C
S
P
is still large enough to
make a hard switching of the lower tree, then the output of the phase detector would
be proportional to RL I0 . As such, the low-pass component of the output of the phase
detector will have the following form
2Q 2Q
vout LPF = RL I0 sin (∆ω) f (t) ≈ RL I0 (∆ω) f (t) (3.37)
ω1 ω1
The linear approximation is valid for 2Q∆ω/ω1 < π/4. As Equation 3.36 suggests,
higher quality factor of the resonant circuit results in larger amplitude of V2 . However,
linear approximation of the frequency response will be violated for large values of
the quality factor and distortion in the baseband data will emerge at the output due
to response nonlinearity. To mitigate this issue, one may decrease the quality factor.
There is another frequency demodulation scheme which is discussed in the PLL section.
Limiter
10kHz
V1
455±8kHz Vout
2nd order
V2
C1
R1 R2
C2
Solution:
(a) Given R1 = R2 = R, and C1 = C2 = C, the transfer function of the Sallen–Key
filter can be written as
v2 1
= (3.39)
v1 (1 + jRCω)2
Then
v2 ω f
∠ = −2tan−1 (RCω) = −2tan−1 = −2tan −1
(3.40)
v1 ω0 f0
v2 ∆f
∠ = −2tan−1 1 + (3.41)
v1 f0
Note that
V2 1 ∆f
≈ for 1 (3.42)
V1 2 f0
v2 π ∆f π ∆f
∠ = −2 + −··· = − − (3.43)
v1 4 2 f0 2 f0
140 Chapter 3. PLL, FM Modulation, and FM Demodulation
Now two signals with large amplitudes and the above phase shift are applied
to a phase detector with a gain of KPD , thus the output voltage of this block for a
sinusoidal modulation will be
π ∆f
vOUT = −KPD + cos (ωmt) (3.44)
2 f0
∆f 8
vout = −KPD cos (ωmt) = −KPD cos (ωmt) (3.45)
f0 455
(b) The 1 kHz component is the main baseband transmitted signal, i.e., ωm =
2π(1000) Hz. If we expand Equation 3.38 for the higher order terms (nonlinear
terms) as well for vOUT , then we reach to
1 ∆f 2 2
π ∆f
vOUT = −KPD + cos (ωmt) − cos (ωmt)
2 f0 2 f0
!
1 ∆f 3 3
+ cos (ωmt) (3.46)
6 f0
The all-pass filter translates this frequency deviation to a specific phase shift,
and consequently to the corresponding voltage at the output of the phase detector.
Moreover, we know that the nonlinear characteristic of phase transfer function
results in harmonic generation of the baseband signal. As such, the output voltage
will be
" ! !
π 1 ∆f 2 1 ∆f 2 ∆f
vOUT = −kPD − + 1+ (3.47)
2 4 f0 8 f0 f0
#
1 ∆f 2 1 ∆f 3
cos (ωmt) − cos (2ωmt) + cos (3ωmt) + · · ·
4 f0 24 f0
Therefore, the amplitudes of the second and the third harmonics are
2
KPD 8
H2 = (3.48a)
4 455
3
KPD 8
H3 = (3.48b)
24 455
3.3 Basics of PLLs and their Application 141
Example 3.2 Consider the FM detector depicted in Figure 3.6. The FM signal
carrier is at 455 kHz with 2 V amplitude. If the frequency deviation is 8 kHz, and
with the assumption of an ideal multiplier with a load resistance of RL = 1 kΩ and
the total bias current of I0 = 1 mA, and assuming CP = 1 nF, CS = 10 pF, and Q = 5,
obtain the detected output signal amplitude.
Solution:
With the given parameters, we can write
Z t
V1 = 2 sin ω0t + ∆ω f (τ)dτ Volts (3.49)
0
For V2 , we have
Z t
CS π 2Q
V2 = Q × 2 sin ω0t + ∆ω f (τ)dτ + − (∆ω) f (t) Volts
CS +CP 0 2 ω0
(3.50)
Z t
π 2Q
V2 = 0.099 sin ω0t + ∆ω f (τ)dτ + − (∆ω) f (t) Volts (3.51)
0 2 ω0
Considering that the Gilbert cell multiplier is driven to its saturation level by both
input signals, the low-pass component of the output becomes
2Q
vout LPF = RL I0 sin (∆ω) f (t) (3.52)
ω1
Now noting that the sinusoidal argument is less than 1 rad, we then reach to
2Q
vout LPF ≈ RL I0 (∆ω) f (t) = 0.176 f (t) Volts (3.53)
ω0
where by substituting the parameters, the output voltage amplitude is 176 mV.
10.7MHz 10.7MHz
+70kHz -70kHz
VIN VOUT
LPF
t VCO t
∆T ∆T ∆T ∆T
change the VCO frequency toward the instantaneous frequency of 10.7 MHz ± 70 kHz.
In other words, the VCO in this loop tries to follow the input frequency. This behavior
is called phase locking and we call this loop as PLL. While the input frequency varies
in the PLL, the loop tries to generate an error voltage to correct the VCO frequency.
This error voltage is proportional to the baseband modulating signal and by this virtue,
the PLL output will be the FM detected signal. Note that if the variations are fast (that
is, faster than the loop bandwidth), the loop would not be capable of following the
input frequency and the loop will not operate properly. Indeed, the PLL is a low-pass
system.
Example 3.3 A student asks whether the PLL is the same as frequency-locked
loop (FLL), i.e., at the steady state, the frequencies will be the same as the phases
are the same. Is he/she right?
Answer:
Yes, in a sense that once the loop is locked the reference and the VCO output
frequencies would be the same but with a constant phase shift existing between
them. But if the loop is not locked the VCO will act as a free-running oscillator.
VOPD
KPD = (3.54)
φin − φout
where in Equation 3.36, KPD is the phase detector gain. We may obtain the transfer
function of the low-pass filter (for a single-pole low-pass filter) easily as
Vout 1
= s (3.55)
VinLPF 1 + ωLPF
VCO
The oscillator in its static mode oscillates at the free-running frequency, ωfr . Here
KVCO is the VCO gain in rad/s per Volts or equivalently Hz V . Depending on the VCO
gain and the input signal, the output frequency changes. We can write the expression
for the instantaneous frequency of the oscillator as
We can reach to time-dependent frequency by taking the derivative of the total phase
which results in
d
Θ (t) = ωfr + KVCOVout (t) = ωosc (t) (3.58)
dt
Taking the Laplace transform of both sides of Equation 3.58 gives
!
1 KVCO
a (s) = s KPD (3.61)
1 + ωLPF s
Then, the closed-loop gain (as a negative feedback loop) which is the transfer function
of interest can be written as
φo a (s)
= (3.62)
φi 1 + f a (s)
Substituting Equation 3.61 into Equation 3.62 and setting the feedback gain f = 1 give
1 KPD KVCO
φo 1+ ω s s
LPF
= (3.63)
φi 1 KPD KVCO
1+ 1+ ω s s
LPF
φo 1
= (3.64)
s2
φi
ωLPF KVCO KPD + KVCOs KPD + 1
Equation 3.64 suggests that the transfer function gain at low frequencies is unity
which means that the loop follows the input phase at the output for low-frequency
variation. However, for a high-frequency input, the gain of the loop will be decreased.
Thus, the transfer function has a low-pass behavior. We can rewrite the transfer
function as
φo 1
= 2 (3.65)
φi s
ωn + Qωs n + 1
The parameter Q in Equation 3.65 has an important effect. If Q is equal to 1/2, the
poles of the loop coincide, if it is greater than 1/2, we will have complex conjugate
poles, and if Q is lower than 1/2, the loop consists of real poles. It is instructive to
know that at ωn , the loop exhibits an overshoot which is illustrated in Figure 3.14.
ORJ݊RXW݊LQ
G% 4G%
G%GHF
ȦQ ORJȦ
Figure 3.14: Overshoot in the frequency response of the PLL near the natural
frequency.
146 Chapter 3. PLL, FM Modulation, and FM Demodulation
Example 3.4 To implement the phase detector, we use the given Gilbert cell.
(a) Find the load resistance and the load capacitance to have a phase-detector gain
of π1 Radian
Volts
.
(b) With the phase detector characteristics depicted in Fig. 3.16, we implement a
PLL as in Figure 3.17. Suppose Rf = 100 Ω, find the value of Cf and the transfer
function of the loop.
(c) If the input frequency suddenly changes from 100 MHz to 100.1 MHz, draw the
control voltage as a function of time.
Assume KVCO = 500 kHz/V, and Q = 1/ (2ζ ) = 0.5.
VDD
Vout
RL CL RL
M3 M4 M5 M6
V2
M1 M2
V1
I0=0.5mA
Vout
+0.5V
∆ɸ
0 π/2 π
-0.5V
%XIIHU
DPSOLILHU
݊LQ 5I .9&2 ݊RXW
*
V
&I
Solution:
(a) When a current completely flows to one side, we have RL I0 = 0.5 V, which
gives the load resistance of 1 kΩ, and consequently (Figure 3.16) the gain of the
phase detector will be 1/π.
As the output of the phase detector should be low pass
1
100 MHz (3.67)
2πRLCL
Let
1
= 5 MHz (3.68)
4πRLCL
Then, CL = 16 pF.
(b) We have the expression for Q as
s
r 1
KPD KVCO (2π × 0.5 MHz)
Q= ⇒ 0.5 = π (3.69)
ωLPF ωLPF
rad
ωLPF = 2π × 6.25 × 105 (3.70)
sec
For the value of the capacitor, we have
Q2
Cf = = 400 pF (3.71)
2πKPD KVCO Rf
φo 1
= 2 (3.72)
φi s
ωn + Qωs n + 1
rad
= 2M sec (3.73)
As Equation 3.74 suggests, the system is critically damped here, and therefore,
it will have the fastest response without overshoot.
(c) The relation between the output frequency and the input frequency can be
written as
fo (s) sφo (s)
= (3.75)
fi (s) sφi (s)
Thus, the output frequency varies with double pole as 1/(s + 2 Mrad/sec)2 , and we
will reach to Figure 3.18.
IR
0+]
0+]
IJ 0UDGV
QV
WIJ
KVCO
φo = vin,control ⇒ sφo = KVCO vin,control (3.76)
s
and for the frequency, we have
Finally, Figure 3.19 depicts how the control voltage varies with time and
reaches to its final value.
9FRQWURO
P9
WIJ
3.3 Basics of PLLs and their Application 149
Example 3.5 In the FM detector circuit depicted in Figure 3.20, the output of
the limiter has an amplitude of 200 mV. With a maximum frequency variation rate
(modulation rate) of 5 MHz, the frequency deviation is 7 MHz. The IF carrier
frequency is at 140 MHz.
ȍ
$
9R
$PS /LPLWHU P9SS
9R
%
ȍ = ȍ
5 Ȝ ȍ
ȍ 5 5 ȍ
5V
RL CL RL
1.5kΩ 7.6pF 1.5kΩ
Q3 Q4 Q5 Q6
400mVp-p
A
RFC 50Ω
Q1 Q2
Vb C∞ C∞
50Ω 1mA
C∞
B
Solution:
(a) In this part, the input signal is attenuated through a π-section resistive attenuator.
As the attenuator is matched at the input and the output, the output voltage will
become
150 Chapter 3. PLL, FM Modulation, and FM Demodulation
R1 k Z0
VB = VA = 0.706VA = 141 mV (3.78)
R2 + R1 k Z0
(b) Here, given the instantaneous FM signal frequency, the quarter-wave transmis-
sion line acts as a 90◦ phase shifter in the following manner.
The instantaneous frequency is
ωi = ω0 + ∆ω f (t) (3.79)
where f (t) is the baseband modulating signal, with unity amplitude. Now, given
the low-pass output circuit of the multiplier and the fact that VA and VB are quite
larger than Vt , the Gilbert Cell acts as an ideal phase detector (recall section 3.2.2),
so its output would be proportional to the phase difference of the in-phase and the
quadrature signals. Considering the cut-off frequency of the output RC circuit as
1
fcut-off = ≈ 7 MHz (3.81)
4πRLCL
Therefore, given the fact that the modulating signal is band limited to 5 MHz, the
output would have the following form
π ∆ω
Vout = I0 RL f (t) = 117 f (t) mV (3.82)
2 ω0
(c) The phase detector works such that it gives a zero DC output for a π/2 phase
shift between the two input signals. Therefore, if the transmission line has a 70◦
phase shift at the center frequency, the DC output would become
!
7π
∆φ 18 2
VDC,out = I0 RL 1 − π = I0 RL 1 − π = I0 RL ≈ 333 mV (3.83)
2 2 9
Some applications mandate high-speed PLLs; however, others may use slow loops. It is
possible to control the loop speed by proper choice of ωn . Moreover, one may change
the bandwidth of the low-pass filter to control the loop bandwidth. Equation 3.66
suggests that lowering the low-pass filter bandwidth results in increase in Q which may
be undesirable and also may make the loop unstable with any additional parasitic pole.
It can be stated that the flat gain is mostly obtained up to ωn frequency. If one increases
the bandwidth of the low-pass filter in order to achieve a fast loop, the bandwidth
will not be extended because of the fact that the poles move farther from each other.
3.3 Basics of PLLs and their Application 151
ωVCO
KVCO
ωfree-running
Vin-DC
Vin, free-running
Thus, it seems that we should increase the gain of the phase detector or the VCO gain
to maintain the quality factor constant. This loop is called Type-I loop, because its
open-loop gain has a pole at the origin (note that the order of the transfer function of
the PLL is always equal to the order of the transfer function of low-pass filter plus
one). Now, consider a tone with 10 MHz frequency is applied at the input of a Type-I
PLL. Moreover, the initial phase difference between the input and the output is 90◦ . If
this input is applied to a Gilbert cell, the output voltage will be zero. If the oscillator is
at its free-running frequency, the loop will be stable. Note that, if the input frequency
changes to 11 MHz, an input voltage must be applied to the VCO to shift its frequency
to 11 MHz. Thus, the phase difference between the input and the output will diverge
from 90◦ , it may be, say, 85◦ . Thus, a PLL is not inherently capable of locking to any
frequency. This phenomenon occurs due to the limited locking and capturing range in
PLLs that is due to transfer function of the phase detector. The consequence of this
phenomenon is that a PLL may not be locked.
Figure 3.22 shows the transfer function of the VCO. It is imperative that the
designer must take into account the voltage range of the phase detector output and the
VCO transfer function to allow the loop to lock.
Example 3.6 Given the initial conditions of the PLL transfer function, how is
that the input and the output frequencies will be equal in steady state?
2π
φo s fo fo 1
= 2π
= = 2 (3.84)
φi s fi
fi s
ωn + Qωs n + 1
Solution:
Bearing in mind that the initial condition must be considered in Laplace transform,
since we describe here the equation about the free-running frequency of the VCO
(Vin−VCO = 0), the phase initial condition is not important here. This point is shown
in the time domain as follows
ϕo (t) − ϕi (t) = cte. (3.85)
1RLV\VLJQDO
1RVLJQDO
,QSXW
VLJQDO
1RUPDO 1RUPDO
VLJQDO VLJQDO
/DUJHVLJQDO
2XWSXW
VLJQDO
Example 3.7 Does the transfer function relating the output phase to the input
phase of the PLL infer unconditional stability, because of the fact that the output
phase reaches to −180◦ at positive infinite frequency?
Solution:
This is the simplified transfer function of the system with two poles; however, due
to nonidealities, the order of the system might be increased. Furthermore, the phase
margin is defined for an open loop, and we write it for the open loop to predict the
closed-loop behavior. Moreover, since the transient response of the loop is of great
importance, we need to take care of the phase margin.
3.4 Further PLL Applications 153
Example 3.8 Consider the given type I PLL in Figure 3.24 with the Gilbert cell
as the phase detector and with the following parameters.
VCO
݊in VPD R Vcont KVCO ݊out
KPD
s
C
(b) To calculate the phase margin, we should find the point where the open-loop
gain reaches unity. Then at that point, we compute the phase. Therefore
1 KVCO
|HOL ( jω)| = 1 ⇒ KPD =1 (3.90)
RC jω + 1 jω
Finally, the phase at this frequency will be ϕ = − π2 − tan−1 (RCω) = −177.1◦ and
the resulting phase margin is 180 − 177.1 = 2.9◦ .
154 Chapter 3. PLL, FM Modulation, and FM Demodulation
(c) Since the given frequency is equal to the free-running frequency of the oscillator,
the phase difference will be 90◦ and the control voltage will be zero.
ϕi − ϕo = 90◦ (3.93)
∆f 10 MHz
∆V × KVCO = ∆ f ⇒ ∆V = = = 50 mV (3.94)
KVCO 200 MHz
V
and the phase difference with respect to the previous case will be
VPD 50 × 10−3
∆ϕ = = = 0.025 rad = 1.43◦ (3.95)
KPD 2
ϕi − ϕo = 90 + ∆ϕ (3.96)
Example 3.9 In the previous example, using the ADS simulation tool, compute
the following. The reference signal at first has a frequency of f1 and then it
experiences a frequency step and goes to a frequency of f2 ,
(a) Draw the control voltage (Vcont ), VPD , Vin , Vout , and fout .
(b) Suppose f1 = 100 MHz, f2 = 110 MHz, and KVCO = 200 MHz/V. Find the final
value of Vcont with respect to its initial value.
V(t)
f1 f2
(c) If the input signal with the frequency of f1 , where f1 is not the free-running
frequency of the oscillator, vanishes, describe qualitatively what happens in the
PLL.
3.4 Further PLL Applications 155
V(t)
f1 f2 f1
Solution:
(a) Figure 3.27 depicts the wanted signals.
(b) We can write
∆f 10 MHz
∆V × KVCO = ∆ f ⇒ ∆V = = = 50 mV (3.97)
KVCO 200 MHz
V
93'
W
9FRQW P9
W
0+]
IRXW
0+]
W
(c) The frequency response of the loop is dependent of its natural frequency, i.e.,
ωn . If this value is large with respect to the input, then the loop will be fast enough
to sense the disappearing of the signal and pushes the VCO to its free-running
frequency. However, if the mentioned disconnection time is small with respect to
the loop time constant, the loop may stand at its current frequency and phase and
the VCO will continue its oscillation properly.
9''
/ /
0 0
0 0
& &
,'&
5
9'&
&
CV(v)
VDC
resonant circuit terminals. The frequency of the oscillation then can be obtained as
s
1 1
f= (3.98)
2π L (CN +CV (v))
where in Equation 3.98, CN is the total capacitance at the output node and Cv (v) is the
nonlinear bias dependent varactor capacitance. The nonlinearity of varactor results
in changing the VCO gain which in fact changes the closed-loop gain and the phase
margin.
In the previous sections, we discussed the frequency demodulation with PLL. Now,
we focus on FM with a PLL. Figure 3.30 illustrates both frequency modulation and
demodulation with PLL.
Figure 3.30 shows the system-level structure of a frequency modulator. We have
seen that by varying the varactor voltage, we are able to make a frequency modulator.
The varactor was the MOS device which was biased in the reverse region. As an
example, consider a 100 mV single-tone input signal in the control voltage of the VCO
with the frequency of 10 Hz as
Moreover, suppose that the VCO is locked to 10.7 MHz. Depending on the bandwidth
or speed of the loop, different outputs can be achieved. If the loop is faster than the
input signal of the oscillator, it doesn’t let the VCO to change its frequency (maintains
the frequency of the loop as stable). On the other hand, for slow loops, the FM will
Demodulator Modulator
VIN VOUT
RS
+
LPF
VMOD
CS
Vout
VCO
VCO
fVCO
10.7MHz
Vin-DC
1.4 1.5 1.6
be materialized. The PLL here plays a main role to hold the intermediate frequency
(the carrier frequency) as constant. From the quantitative analysis point of view, we
remember that loop’s f−3dB is selected near ωn to have complex poles with proper
settling time. Now assume that the natural frequency of the loop is 500 Hz. Thus, the
loop is fast enough not to let the frequency change. Now, if the input signal to the VCO
changes its frequency to 5 kHz, the VCO changes its frequency with a rate of 5 kHz.
The frequency deviation in the oscillator is merely dependent on the variations of its
control voltage. It is clear that the larger-signal input to the VCO will result in more
frequency deviation from the center frequency of the oscillator. Figure 3.31 depicts the
characteristics of the oscillator for this example.
In an ordinary PLL, the output follows the input to find the same frequency.
However, in FM with PLL, the loop resists against the carrier frequency variation. In
fact, the loop has an output with the average frequency of 10.7 MHz and will find
a frequency deviation corresponding to the input signal. It can be stated that in the
frequency modulator, the loop should be designed as a slow loop, and in the frequency
demodulator, the loop should be designed as a fast loop. Thus, the lower limit in
frequency modulator is ωn and the upper limit is specified by the low-pass filter for the
modulating signal.
Example 3.10 Is is possible to feed the baseband signal to the VCO for the sake
of FM generation without a PLL?
Answer:
Although an FM modulator with a simple VCO is conceivable, practically it is
not possible, because of the requirement for the carrier frequency stability. The
frequency stability of the PLL is then necessary for correct operation of the FM
generation which is guaranteed by means of the negative feedback in the PLL loop.
Moreover, PLL shapes the phase noise of the oscillator which is of great importance
as well.
RS
+
VMOD
Vout
÷N VCO
÷M
Figure 3.32 is the usual frequency synthesizer which is used to generate a frequency-
modulated signal. The divider shown in Figure 3.32 is a digital counter which after M
input pulses, generates one pulse. Note in any counter, the value of M can be selected
digitally. In the steady-state condition, both the inputs of the phase detector will have
the same frequency and as a result, we will have
fCrystal fVCO
= (3.100)
N M
Suppose the crystal frequency is equal to 10.7 MHz, then for the VCO frequency, one
may obtain
M M
fVCO = fCrystal = 10.7 MHz (3.101)
N N
Moreover, assume that for the input divider, we have N = 107. Thus, the comparison
frequency will be equal to 100 kHz. Now, assume that M = 9000. As a result, the output
signal will be at 900 MHz and the channel spacing could be 100 kHz. The channel
selection can be achieved by changing M, and thus 10.7NMHz will be the minimum
channel step. Assuming an input sinusoidal signal with 100 mV for the modulation
signal, for the frequency of the VCO, we will have
M
fVCO = fCrystal + 100 mV × KVCO sin (ωBBt) (3.102)
N
M is changed by the digital circuitry, and therefore one may hop from one channel to
another. With respect to different standards, we can change the comparison frequency
to change the channel spacing. In the high-frequency applications (e.g., higher than
5 GHz), we should break the divider into several stages and design a special counter
for the first stage which operates at high frequency.
Till now, we have learned how to demodulate a frequency-modulated signal by
a quadrature resonator or a PLL. Suppose the input signal frequency to the PLL is
10.7 MHz ± 70 kHz (in other words, the frequency deviation is 70 kHz), therefore the
VCO follows the input frequency variations and the output of the phase detector through
the low-pass filter gives in the detected FM baseband. However, in the quadrature FM
detector, if the transmitted signal carrier frequency is changed, the detector could not
detect thoroughly the input FM because the phase shift in quadrature component will
be no longer about 90◦ and the detector would not perform correctly.
We have also shown that, using a PLL, we are able to generate a frequency-
modulated signal which is shown in Figure 3.33. As stated earlier, the bandwidth of
the FM signal at the PLL output in Figure 3.33 is
BW = 2 ( fdev + fm ) (3.103)
where fdev is the maximum frequency deviation and fm is the maximum frequency of
the baseband signal. However, the bandwidth of the PLLs is far less than the above
bandwidth. The frequency deviation is proportional to the amplitude of the modulating
signal. As discussed earlier, the correct operation of the frequency modulator has two
160 Chapter 3. PLL, FM Modulation, and FM Demodulation
56
902'
5
$FWLYH
FLUFXLW
& / & 5
/3) 9&2
margins which are specified by the natural frequency of the loop and the maximum
frequency component of the modulating signal. In fact, the time constant of the loop
characterizes the upper margin of the PLL. For increasing the time constant, making
the loop slow, we can increase the capacitance in the loop filter.
Now, we derive equations for the dynamic behavior of the synthesizer in Figure 3.32.
We stated that in the steady-state condition, both inputs of the phase detector will have
the same frequency. It can be asserted that the phase detector is somehow a frequency
detector as well and we can employ frequency modulator system as a phase modulator
block too. Now, if we write the relation between the output phase and the input phase
in Figure 3.32, we reach to
−1
φo fo
s
1 + ωLPF KPD KVCO
s
= = −1 (3.104)
fi
φi
1 + 1 + ωLPF s
KPD KVCO 1
s M
Thus, Equation 3.104 gives the transfer function of the frequency synthesizer. One of
the important parameters in this loop is the transition time to shift from one channel
frequency to another channel frequency which can be calculated through the inverse
Laplace transform of Equation 3.104 which yields the settling time as
4
TS = (3.105)
ζ ωn
Note that the settling time is defined as the lapse of time required for the output
frequency to reach 98% of its final value. Here ζ is the damping factor and it is
expressed by
1
ζ= (3.106)
2Q
Now, suppose the oscillator in Figure 3.32 has a frequency equal to 900 MHz and the
channel spacing is 30 kHz. We can obtain channel spacing as follows
fCrystal
Channel Spacing = (3.107)
N
3.5 Advanced Topic: PLL Type II 161
Here, N determines the channel spacing. Thus, we now have implemented a frequency
synthesizer which is used for frequency generation for both the receiver and the
transmitter. Moreover, by putting a baseband signal in the control voltage of the latter,
we will have a frequency modulator. As an example, for a crystal oscillator of 15 MHz
frequency, one may obtain the value of N as
15000
N= = 500 (3.108)
30
To change the channel frequency, we are able to change the value of M. Another usual
method for frequency synthesis is the direct digital synthesis (DDS) which is very
precise with the precision of hundredth of hertz (at IF frequency). The DDS-based
design is out of the scope of this text.
Nowadays, the FM is not used in high-speed and high-performance transceivers.
It is mainly used in commercial broadcast systems which depend on great number
of conventional FM receivers. However, digital modulations such as M − QAM and
QPSK are common in data communication which we discuss in the following chapters.
VDD
D
Q QA A
A CLK
B
Reset AND
QA
B CLK
Q QB QB
D
VDD
Charge pump
VDD
I1
Up S1 Vo
Down S2
C
I2
Figure 3.35: A typical charge-pump circuit using two current sources and two
switches.
7
9''
$
,ϭ ǻ݊
%
$ 6ϭ 9FRQW
3)' 4$
% 6
&
,
9FRQW , ǻ݊
Ϯʌ&
W
Figure 3.36: Operation of the charge-pump circuit under the excitation of the
PFD.
3.5 Advanced Topic: PLL Type II 163
Figure 3.36 suggests that when the switch is on, the capacitor is charged linearly
and the circuit can be assumed as an integrator and when the switch is off the capacitor
holds its value. The output voltage increment in Figure 3.36 can be approximated as
∆ϕ I
∆Vcont = T (3.109)
2π C
Equation 3.109 can be rewritten for the control voltage of the oscillator as
∆ϕ I
Vcont (t) = tu (t) (3.110)
2π C
By taking the Laplace transform of Equation 3.110, we reach to
Vcont (s) I 1
= (3.111)
∆ϕ 2πC s
Equation 3.111 shows the integration behavior of the circuit explicitly. Finally, by
placing the charge-pump circuit subsequent to the PFD, and applying a unity feedback,
the type II PLL can be achieved as in Figure 3.37.
The reason that we call this architecture type-II is that it has two poles at the
origin in the open-loop transfer function (one for the charge pump and another for
the VCO). The two poles at the origin make the instability of great concern. Thus,
for the stability issues, we place a series resistor with the capacitor and rewrite the
charge-pump equation as (this brings a zero in the open-loop as well as the closed-loop
transfer function)
Vcont I 1
(s) = +R (3.112)
∆ϕ 2π Cs
VDD
I1
ɸi S1 Vcont
PFD VCO ɸo
S2
C
I2 R
Figure 3.37: Type-II PLL block diagram including a PFD, a charge pump, and
a VCO.
164 Chapter 3. PLL, FM Modulation, and FM Demodulation
Thus, we can write the overall transfer function for type-II PLL as
IKVCO ω 2 1 + 2ξ s
ϕo 2πC (RCs + 1) n ωn
H (s) = = 2 I I
= 2 (3.113)
ϕi s + 2π KVCO Rs + 2πC KVCO s + 2ξ ωn s + ωn2
Furthermore, the poles and the zero for the transfer function, H(s), are
p
sp1,2 = −ξ ± ξ 2 − 1 ωn (3.115a)
−ωn 1
sZ = =− (3.115b)
2ξ RC
As Equation 3.114 suggests, to mitigate the spur level, we can increase the value of
C, and therefore ζ will be increased which now does not pose any problem for the
instability. Thus, the drawbacks of the type-I PLL are now resolved at the cost of lower
phase margin and consideration for stability due to increased order of the transfer
function. To increase the locking speed, one should increase ωn , and therefore, IKVCO
should be increased, or C could be decreased. Regarding the stability check of the
type-II PLL, further reading in the given references is recommended.
3.6 Conclusion
In this chapter, the general configuration of the PLLs was studied. Care should be
taken that in a PLL, the parameter of the study whose stability and response should be
considered is the phase (and consequently, the frequency), so here we are considering
the frequency response of the phase (or the frequency) in the loop. The phase detector
was one of the main components of the PLL whose implementation using a Gilbert
cell or an XOR was introduced. FM using a varactor-tuned oscillator was introduced
alongside an FM demodulator using a quadrature resonator. The FM demodulation
is possible using a sufficiently high-speed PLL. This concept was introduced as well.
FM is possible using a low-speed PLL whose concept was described in this chapter.
Frequency synthesizers are one of the basic building blocks of the modern transceivers.
The basic structure of a frequency synthesizer using a crystal oscillator, a frequency
divider, and a PLL including a second frequency divider was introduced as well. Type I
PLLs are based on a phase detector, a low-pass filter, and a VCO. This type of PLL
suffers from the problem of instability, and limited locking range. Type II PLLs were
3.7 References and Further Reading 165
introduced to mitigate the problem of instability and the locking range. The type II
PLL is based on a phase-frequency-detector, a charge pump, and a VCO. In sum,
the building blocks described in this chapter can be used as frequency modulators,
frequency demodulators, synthesizers, and eventually phase modulators.
3.8 Problems
Problem 3.1 Figure 3.38 depicts a simplified frequency synthesizer.
fref R KV fout(900MHz)
KP
s
C
÷M
30KHz
900MHz
In the transfer function of the loop, the value of ζ is taken as 0.707 and ωn =
500 rad/sec, and KP = 10 V/rad. q q
ωLPF KV KP 1 MωLPF
1. First show that in this loop ωn = M , ζ = 2 KV KP .
2. For 30 kHz reference frequency, design the synthesizer for the channel spacing
of 30 kHz and a center frequency of 900 MHz. (Find the divider’s modulus M,
the low-pass filter’s RC time constant, and the VCO gain, KV ).
3. Find the settling time of the loop when it hops from the current channel to the
adjacent channel.
4. If we replace the phase detector with a bipolar Gilbert cell, find the value of RL
for a bias current of 5 mA to obtain KP = 10 V/rad.
ĭref R KV ĭout
KP + s
fref= C
15MHz
Xin
÷60
2. If the signal xin is injected to the input of the VCO through RS , determine
the minimum and maximum frequency of the baseband input. Suppose that
R = 10 kΩ, C = 159.2 nF, RS = 100 kΩ, and the average capacitance seen
through the VCO is Cin,0 = 160 pF. You may use the equivalent circuit shown
in Figure 3.40 for this purpose.
3. If R1 = 400 Ω, and KV = 2π × 100 kHz/V determine the required KP and conse-
quently the tail current of the Gilbert cell phase detector.
$FWLYH
FLUFXLW
[LQ
*LOOEHUW3'
5V
5 /3)
9&2
5
*
&
&LQ
9DUDFWRU
Figure 3.40: The equivalent circuit of the part of the PLL used as the FM
modulator.
ωLPF
Vout
÷N VCO
15MHz
N=15
÷M
M=1000
Problem 3.4 FM Modulator; In the MOS oscillator stage depicted in Figure 3.42,
the right-hand section acts as a variable reactance which loads the left-hand oscillator
1
stage. Here, assume that r Cω 0
. Determine an expression for the variable reactance
seen through the right-hand section and from there obtain an expression for the fre-
quency of oscillations (the carrier frequency and the frequency deviation) in terms of
168 Chapter 3. PLL, FM Modulation, and FM Demodulation
the circuit parameters. Assume that both MOS transistors operate in the square-law
active region. Secondly, write an expression for the oscillation condition which deter-
mines the amplitude of oscillations. Here, f (t) is the low frequency baseband signal
varying between +1 and −1.
M12 VDD
Variable
reactance
L2 L1 R1 C1
M1 M2
+
Vinf(t)
r -
ID C∞
VGS0
-VDD
IDS=K(VGS-VTH)2
D Vout
+
V1cosω0t
C R L
+
vScosωSt
of, say, 700 mV and shows a finite turn on resistance. We know the on-resistance of
the diode is equal to
VT
ron = (4.1)
ID
Normally, the turn-on resistance of the diode is in order of the few ohms which could
be considered as a short circuit compared to the load resistance (R). As such, once
the diode is on, a whole RF voltage would appear at the output. Once the diode is
turned off (has a large series impedance) in the negative half cycle of the LO signal, the
output voltage goes to zero. As such, the input RF signal is sampled at the rate of the
LO signal. The output voltage can be expressed as
where S(ω0t) is a square-wave signal that toggles between one and zero with the period
of the LO. Its Fourier expansion is expressed in Equation 4.40. The small-signal output
waveform is shown in Figure 4.2. It is obvious that within the right-hand product of
Equation 4.2, there exists the sum and difference frequency components of the RF and
the LO terms. As such, if the RF is at the input, the difference component gives in the
IF signal and if the IF was at the input, the sum component would give in the RF signal.
For now, we have shown that the large-signal input makes diode to be on and off
and when the diode is on, the input small signal appears at the output and when the
diode is off, there would be no signal at the output. Moreover, by virtue of the tuned
circuit, the desired frequency component of the signal would appear at the output. In
the next section, we delve into the nonlinear transconductance which is approximated
by a polynomial expansion.
vScosωSt
1.5V
0.7V
V1cosω0t t
Vout t
RS i=av+bv2+cv3
+ +
VLcosωLt L C R
Rin v i
+
VRcosωRt
complex function such as an exponential one. In radio communication, the weak signal
is received by the antenna which is noted in Figure 4.3 as (VR cos ωRt) and then this
signal is mixed with the local oscillator signal noted as (VL cos ωLt). Therefore, one
may express the output current as
Our objective is to find the product term of RF and LO frequencies in Equation 4.3.
We can expand Equation 4.3 to arrive at Equation 4.4:
i = aVR cos(ωRt) + aVL cos(ωLt) + bVR 2 cos2 (ωRt) + bVL 2 cos(ωLt) (4.4)
3 3 3 3
+ 2bVRVL cos(ωLt) cos(ωRt) + cVR cos (ωRt) + cVL cos (ωLt)
+ 3cVR 2 cos2 (ωRt)VL cos(ωLt) + 3cVR cos(ωRt)VL 2 cos2 (ωLt)
Each nonlinear circuit is capable of receiving both large and small signals, and by
virtue of its nonlinearity generates the harmonics of the inputs and their products. We
can define each component of Equation 4.4 as “RF,” “LO” themselves, and “RF 2nd
harmonic and a DC component,” “LO 2nd harmonic and a DC component,” “desired
component of IF,” “3rd harmonic of RF,” and finally “3rd harmonic of LO.” With the
following trigonometric equations
1 + cos (2ωt)
cos2 (ωt) = (4.5)
2
3 1
cos3 (ωt) = cos (ωt) + cos (3ωt) (4.6)
4 4
In real design, however, the large signal is the signal of local oscillator which can
degrade the performance of the circuit due to nonlinear characteristic of diodes. More-
over, this signal can leak to other points of the circuit through the supply voltage line
and the ground line, and cause undesirable effects. This leaked signal upon a nonlinear
element can generate unwanted harmonics and mixing products. Finally, the main
drawback of a nonlinear system is the handling of strong interferes and intermodulation
products. This unfavorable mixing occurs in any nonlinear circuit with large-signal
input. Tuning circuit may be useful to mitigate the effect of harmonic generation. For
instance, if the LO frequency resides at 945 MHz and the RF frequency is at 900 MHz
(as in the GSM case), by tuning the resonant circuit at 45 MHz, we can suppress the
unwanted mixing products.
172 Chapter 4. Mixers
Now, we can expand Equation 4.8 to obtain all the harmonic at the output. Until
now, we have carried out equations for the output harmonics of a nonlinear circuit.
Another important issue in a nonlinear amplifier is named as intermodulation (IM).
Our IM of interest is IM3 which is the intermodulation product caused by third-order
nonlinearity. Regarding two inputs as v1 = Vf cos (ωft) and v2 = Vm cos (ωmt) as to
adjacent channels, with respect to Equation 4.4, we then reach to
VO =aVm cos(ωmt) + aVf cos(ωft) + bVm2 cos2 (ωmt) + bVf2 cos2 (ωft) (4.9)
+ 2bVfVm cos(ωft) cos(ωmt) + cVm3 cos3 (ωmt) + cVf3 cos3 (ωft)
+ 3cVm2Vf cos2 (ωmt) cos(ωft) + 3cVmVf2 cos(ωmt)cos2 (ωft)
3 3
VIM = cVm2Vf cos((2ωm − ωf )t) + cVmVf2 cos((2ωf − ωm )t) (4.10)
4 4
Figure 4.4 depicts the signal spectra at the input and the output of the nonlinear
amplifier.
As Figure 4.4 suggests by the virtue of nonlinearity in the amplifier, different
mixing products of the two input signals are generated at the output. However, in this
derivation, we have merely taken into account a polynomial of third order. Magnitude
of each component in Figure 4.4 can be easily computed by Equation 4.9. The
green component in Figure 4.4 is called the IM product of third order, because this
4.2 Third Order Intermodulation Concept in a Nonlinear Amplifier 173
1RQOLQHDU
FLUFXLW
9L 9PFRVȦPW9IFRVȦIW E9LF9L
9R D9L
,QSXW 2XWSXW
VSHFWUXP VSHFWUXP
ȦPȦI
ȦIȦP
ȦPȦI
ȦIȦP
ȦIȦP ȦIȦP
Ȧ Ȧ
ȦP ȦI '& ȦP ȦI ȦP ȦI ȦP ȦI
Figure 4.4: Representation of mixing products for two input adjacent channels.
term is generated due to the cubic term of the polynomial. This component can be
troublesome in wide-band receivers and we then linearize the amplifier to mitigate
this effect. As an example, suppose we have two adjacent channels with the frequency
of ωf = 2π × 100.2 MHz and ωm = 2π × 100 MHz. Thus, IM3 components reside at
2ωf − ωm = 2π × 100.4 MHz and 2ωm − ωf = 2π × 99.8 MHz. As each channel is
normally modulated by a random signal, the intermodulation products (IM3) could be
considered as a random noise for either of the channels. Thus, this might be a drawback
in receivers which can degrade signal-to-noise ratio (SNR) of the alternative channel.
Assuming the magnitude of adjacent channel equal to V , Equation 4.9 suggests that the
IM3 competent grows by V 3 and each channel power grows by V . This is an important
point which exacerbates more the SNR. Note, if the power of each channel is added
by 1 dB, IM3 component power will be added by 3 dB. This concern is mitigated by
linearizing nonlinear circuit.
2XWSXW
G%P 6ORSH
RI 6ORSH
2,3 RI
G% 2XWSXW
FRPSUHVVLRQ
,QSXW
G%P
3LQG% ,,3
approximately equal to one (or 10 dB/dec). The lower nonlinear curve shows the
evolution of the IM products level with respect to the input level. Its slope at the lower
values of the input is about three times of that of the main output (30 dB/dec). Both
of these curves saturate (experience a decrease in their respective slopes) at the high
levels of the input signals.
If one draws the tangents at the two curves at lower signal levels and extends them
far enough towards the higher levels, the two lines would intersect at a point which we
call the third input-intercept-point (IIP3 on the abscissa). Moreover, the output point is
called oip3 . Figure 4.5 shows a real compression of the output signal which is denoted
by the green line. In fact, the IIP3 point is a practical indication of the nonlinearity
of the amplifier. The higher it is, the more linear is the amplifier. The lower it is, the
more nonlinear is the amplifier. Another point of interest is the saturation point of the
amplifier and that point is where the difference between the linear input/output (tangent
line) characteristic and the nonlinear (the real) input/output characteristic comes to
1 dB difference value, is called the compression point. It is another indication of the
linearity of the amplifier. The higher the compression point, the more linear is the
amplifier. Regarding the compression point refer to (equ compression),
3
vmo = aVm + cVm 3 cos (ωmt) (4.11)
4
3
vfo = aVf + cVf 3 cos (ωft) (4.12)
4
RS i=av+bv2+cv3+dv4
+ +
Vmcosωmt
+ i L C R
Vfcosωft Ci v
+
VLcosωLt
-
Now, if we just look at the low-pass signal components of Equation 4.18, we can obtain
,QSXW 2XWSXW
VSHFWUXP VSHFWUXP
I0+] I0+]
0+]0+] .+].+]
5)
'HVLUHG 0+]
,QWHUIHUHU 0+]
/2 0+] /2 0+]
,QWHUIHUHU 0+]
î
2XWSXW
î
î
VSHFWUXP
I0+]
Figure 4.9: The spectrum of signals for the mixer of Fig. 4.8.
The frequency components in Equation 4.19 are IM3 which should be taken into
account from linearity perspective. Whenever we record the input–output characteristic
of a linear system, we reach to a line with the slope of one which shows the small-signal
constant gain. In other words, if the input grows with just 1 dB, the output will be added
by the same value. However, in nonlinear systems, IM3 component will experience
3 dB growth with 1 dB input increase. For a highly linear mixer, the IIP3 value is
high. The problem that may arise is in the fading case of the desired signal and the
presence of high-level adjacent interfering (blocker) channels. The IM3 components of
the strong adjacent channels might fall within the reception bandwidth of the receiver.
This may be troublesome in radio systems. In the real world, however, this issue can
be alleviated by frequency hopping and the use of error-correcting codes. The presence
of a strong blocker (interferer) signal in a nonlinear mixer is a challenge. One way to
handle this challenge is to linearize the mixer.
Another IM component is IM5 that increases by 50 dB/dec of input increase and
has emerged by virtue of a term with the sixth order in the nonlinear model of the
transconductance. As stated earlier, IIP3 is the parameter which gives a measure of
linearity in a system, thus we intend to find an easy method to compute it through
input/output measurement. It can be proved that this value can be written as
∆dB
IIP3dBm = inputdBm + (4.20)
2
where in Equation 4.20, ∆dB is the difference between lines of output signal (slope one)
and output intermodulation (slope three), please refer to Figure 4.11. This equation is
proved in the next subsection.
R2 VLcos(ȦLt)
V1=Vmcos(Ȧmt) R1
-
Output
V2=Vfcos(Ȧft) R1 + VA
VA=-(R2/R1)(V1+V2)
2XWSXW
G%P ǻ
2,3
ǻ
1RLVH
OHYHO ,QSXW
G%P
,,3
,QSXW
OHYHO
Figure 4.11: Interpolation of the output signal and the intermodulation curves
to obtain third-order intercept point (IIP3 ).
Our goal is to derive a simple equation for IIP3 . To obtain IIP3 , first of all choose
a point which is in the low input power region for both lines. This is done for a
better approximation of the slopes of the tangents to those curves. Then, by a simple
subtraction of the dB levels recorded on those two lines, divide it by two, and adding
this value to the selected point operating value, we reach to IIP3 . That is
∆P (dB)
IIP3 (dBm) = Pin (dBm) + (4.21)
2
where ∆P is the signal to IM ratio in dB.
IIP3 could be roughly estimated at 7 dBm for a silicon diode mixer in the 50 Ω
system. In a silicon bipolar transistor Gilbert cell, it varies between −20 dBm and
−12 dBm at the input, and for its MOS counterpart, this value is in the range of
−15 dBm to −5 dBm. For applications which necessitate highly linear mixers, IIP3
can be up to 14 dBm. High IIP3 mixer is of great importance in radio systems. In
nonlinear systems and in the presence of interfering channels, signal detection is
somehow tough. In reality, the lines in Figure 4.11 never reach to one another due
to the compression phenomenon; however, the tangent lines give us the measure of
nonlinearity. If one decreases the level of the input signal, such that the desired output
component goes under the noise floor, that point determines the mixer sensitivity. On
the other hand, if we increase our input signal such that the output goes beyond the
compression point, and the resulting distortion in the signal causes error in the received
4.3 Basic Concept of Third-Order IM in a Basic Mixer 179
bits, that point is considered as the saturation point of the mixer. The difference in dB
between those mentioned levels defines the dynamic range of the mixer. To check the
accuracy of ones measurement, one may increase the signal by 1 dB and check the
IM3 component to increase by 3 dB. Another important point in Figure 4.11 is the 1 dB
compression point which is noted by p1 dB . Due to nonlinearity, the gain of the mixer
will be decreased, and the point where the gain drops by 1 dB is of great importance.
In practical system design, we usually work at a back-off (at a level 6 dB–10 dB lower
than the compression point to assure the required linearity) of roughly between 6 dB
and 10 dB with respect to compression point to prevent compression. In modern
applications, we need new techniques to manipulate IM component for better signal
detection. Figure 4.12 depicts a conventional receiver example.
We can also use a mixer to upconvert the signal, in a transmitter which is shown in
Figure 4.13.
In transmitters, both the LO and IF signal are large signals. 900 MHz band-
pass filter is placed to attenuate the other component of mixing residing at 990 MHz
LO2=945 MHz
LO,I=45 MHz
LO2=945 MHz
LO,Q=45 MHz
0 dB
200 30dB
KHz 60dB
400 KHz
Figure 4.14: Wide-band spectrum standard for GSM and ACPR effect.
180 Chapter 4. Mixers
VCC
-
R1 C R L Vo
+
CB
Q2
+
VScosωSt R2
RE CE
VLcosωLt
+
Here, the total base–emitter voltage consists of a DC voltage, a local oscillator voltage,
and an input signal voltage:
By substituting the corresponding values of DC, LO, and input signal voltages in
Equation 4.22, we obtain
ie (t) = IES eqVBEQ /kT eq(VS cos(ωS t))/kT eq(VL cos(ωL t))/kT (4.24)
Here, In (x) or Im (y) are modified Bessel functions of the first kind which exponentially
increase with respect to their argument. It is noteworthy that I0 (x) tends to unity when
its argument tends to zero. In (x) for n ≥ 0 tends to zero when its argument approaches
zero. Furthermore, I1 (x) ≈ x/2 for x < 1. It should be added that
In+1 (x)
< 1, f or x > 0, n ≥ 0 (4.26)
In (x)
Equation 4.25 can be simplified to
" #
∞
qVBEQ /kT Im (y)
ic (t) = αIES e I0 (y) I0 (x) 1 + 2 ∑ cos (mωSt)
m=1 I0 (y)
" #
∞
In (x)
1+2 ∑ cos (nωLt) (4.27)
n=1 I0 (x)
Considering small-signal input and a large-signal LO, and given the fact that I0 (y) ∼
= 1,
I1 (y) y
I (y) ≈ 2 , one can rewrite the collector current expression as
0
I1 (x) I1 (x)
ic (t) = αIE0 1 + y cos (ωSt) + 2 cos (ωLt) + 2y cos (ωSt) cos (ωLt) + · · ·
I0 (x) I0 (x)
(4.31)
By separating the different components of the collector current, one can deduce from
Equation 4.31 that each component of the collector current appears through a certain
transconductance as it is followed.
The input signal frequency component would appear in the collector through a
small-signal transconductance, namely, gm :
IS = αIE0 y = gm vS (4.32)
The local oscillator frequency component would appear in the collector through a
large-signal transconductance, namely, Gm (x):
2I1 (x) 2I1 (x)
IL = αIE0 = gm vL = Gm (x) vL (4.33)
I0 (x) xI0 (x)
2I1(x)/(x.I0(x))
Gm(x)/gm
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0 x
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
I1(x)/I0(x)
gc/gm
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0 x
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
5/ 5/ 5/ 5/ 5/
9RXW
9RXW 9RXW
4 4
9/2 4 4 4 4
9/2 4
9/2
95) 4
4 4
95) 4 95)
,((
D E F
5/ 5/ 5/ 5/ 5/
9RXW
9RXW 9RXW
0 0
0 0 0 0
9/2 0 9/2
9/2
95) 0
0 0
95) 0 95)
,((
G H I
Figure 4.18: Different mixer circuit topologies, (a) bipolar unbalanced, (b) bipo-
lar balanced, (c) bipolar double-balanced, (d) MOS unbalanced, (e) MOS
balanced, and (f) MOS double-balanced.
4.5 Mixer types Based on Switching Circuits 185
Unbalanced Mixer
One may obtain the output signal of mixers (a) and (d), considering a nonlinear power
series transconductance for the lower transistor switched by the LO driven upper
transistor in Figure 4.18 as
2
1 2 2
Vout = a + bVRF + cVRF + · · · + cos (ωLOt) − cos (3ωLOt) + · · · (4.41)
2 π 3π
Equation 4.41 is written using Equation 4.40. To simplify the operation of the mixers,
we can state that Q2 turns on and off by the LO signal. This implies that when this
transistor is on, the RF signal appears at the output; otherwise, the output is tied to the
supply voltage. Thus, we can assume that the RF signal is multiplied by a square wave
with the amplitude of 0 and 1 by the LO period. Therefore, we can attain a new set of
coefficients as
1
b0 = (4.42a)
2
an
bn = (4.42b)
2
Equation 4.41 suggests that there will be lots of mixing products at the output of the
mixer. Thus, we usually employ a low-pass filter at the output to suppress the unwanted
products. Moreover, note that the leakage of RF and LO signals to the output has come
from the DC within the parenthesis terms in Equation 4.42.
Single-Balanced Mixer
We can derive the output signal of mixers
(b) and (e) in Figure 4.18 as for Equation
4.43,
4 4
Vout = a + bVRF + cVRF 2 + · · ·
cos (ωLOt) − cos (3ωLOt) + · · · (4.43)
π 3π
The important point in Equation 4.43 is the effect of differential circuit on the Fourier
series of the LO frequency. It seems that the RF signal now is multiplied by a square
wave with the alternative amplitudes 1 and −1. Thus, no DC component at LO Fourier
series coefficients suggests no RF feedthrough at the output.
186 Chapter 4. Mixers
Double-Balanced Mixer
Finally, the output of mixers (c) and (f) in Figure 4.18 can be written as
3 4 4
Vout = (bVRF + dVRF + · · ·) cos (ωLOt) − cos (3ωLOt) + · · · (4.44)
π 3π
Equation 4.44 introduces no DC components at both LO and RF sides, thus the concept
of double-balanced mixer which doesn’t permit these signals to appear at the output
mixer is obvious. In fact, with this powerful analysis, we are able to compute any
mixing product gain and moreover understand the port-to-port leakages. Nonetheless,
with inevitable mismatches and offset voltages, a finite leakage signals would be
present at the output of the mixer. Today, MOS process offers very fast switches due to
lower capacitances and on-resistances which can operate for high frequencies. One
of the most important specifications of mixers is their linearity issue which has come
from the nonlinear transconductance of the input transistor which converts the input
RF voltage to the current that passes through the switch loads. LO signal applied to
the other transistors just turns them on and off and roughly doesn’t affect the linearity
issues. Another type of mixer which is called a passive switching mixer is shown in
Figure 4.19.
These circuits manifest better linearity because of no transconductance device
between the switch and the load. In other words, the signal itself is chopped by means
of switches and reaches the output. Figure 4.20 depicts a differential implementation
of a passive mixer which is somehow alike active ones without transconductance.
9/2
9/2 5 5
56 56
9,) 9,)
9/2 5 5
95) 95)
9/2
VLO+
VLO+ VLO-
VRF -
VLO- VRF-
R R
VRF+ VLO+ VRF+
VLO- VLO+
VLO-
Equation 4.46 shows that input signal does not appear at the output and the LO leakage
is equal to 4/πRL Ibias . Finally, the double-balanced mixer output signal for mixers
(c) and (f) can be calculated as
4 4 4
Vout = gm RLVin cos (ωRt) cos(ωLt) − cos(3ωLt) + cos(5ωLt) − · · ·
π 3π 5π
(4.47)
where it shows there is no leakage to the output. However, with the definition of
conversion gain, i.e., the gain from IF signal to RF can be carried out as
Vout (IF) 1
= gm RL for unbalanced (4.48a)
Vin (RF) π
Vout (IF) 2
= gm RL for single-balanced (4.48b)
Vin (RF) π
Vout (IF) 2
= gm RL for double-balanced (4.48c)
Vin (RF) π
Note that the coefficient 4/π has come from the Fourier series expansion and 1/2 is
due to one of the sum or difference components obtained out of the multiplication
of cosines. Nowadays, double-balanced mixers are more frequently applied due to
suppression of port-to-port leakages. MOS devices present proper switches for mixing
purposes; however, their quadratic I −V characteristics are such that for a given bias
current, MOS devices have lower transconductance than their bipolar counterparts.
Moreover, note that their output impedance is lower than those of bipolar devices which
is not a merit. It is instructive to note that the main parameter in mixers is their linearity
issue rather than their conversion gain. Moreover, to alleviate the linearity issue, we
should linearize the input active device, because the upper side in the aforementioned
188 Chapter 4. Mixers
mixers is just switches. Figure 4.21 shows a bipolar unbalanced mixer with a tuned
circuit load.
In Figure 4.21, LO signal is connected to the base of Q1 and RF signal is applied
to Q2 . LO signal is large and might have the amplitude of a few hundred millivolts or
more and the RF signal is small. Transistor Q1 will roughly be on and off within each
LO period. When this device is on, it let the current flow to reach the resonant load
and the output voltage appears across the tuned circuit load. However, when Q1 is off,
the current passing through the Q2 collector is nearly zero and the output will be tied
to VCC . Figure 4.22 illustrates the concept of mixing in the mixer in Figure 4.21.
In each cycle, the following happens:
1. Q1 is off (negative half cycle of LO): in this case IC = 0.
2. Q1 is on (positive half cycle of LO): in this case IC = IE0 + gmVR , where IE0 is
VBB2 −VBEQ
IE0 = (1 + β ) (4.49)
R2
Then, in sum, the collector current of Q2 can be expressed as
iC (t) ' [IE0 + gmVR cos (ωRt)] S (ωLt) (4.50)
VCC
C R L
Vo
R2
Q2
+
VRcosωRt
VBB2
R1
Q1
+
VLcosωLt
VBB1
TRF
where S (ωLt) is a monopolar square wave varying between zero and one at the rate of
LO. Then
qIE0
iC (t) ' IE0 + VR cos (ωRt) S (ωLt)
kT
qIE0 1 2
= IE0 + VR cos (ωRt) + cos (ωLt)
kT 2 π
2 2
− cos (3ωLt) + cos (5ωLt) + · · · (4.51)
3π 5π
Finally, if the RLC circuit is tuned to the difference frequency, the output AC voltage
becomes
RL qIE0
vout ' VR cos ((ωR − ωL )t) (4.52)
π kT
In another mode of operation, a similar circuit topology can be used as an upconverting
mixer. Here a bypass capacitor CE is used between the Q2 emitter and the ground
(Figure 4.23). This capacitor should be sufficiently large to be short at the LO frequency
and adequately small to be open at the IF frequency. As such, the transistor Q1 acts as
a time-varying current source biasing Q2 at the rate of IF. Here, we have
VCC
C R L
Vo
R2
Q2
+
VLcosωLt
VBB2
Q1 CE
+
VIFcosωIFt
VBB1 RE
2I1 (x)
Gm (x) = gm (4.54)
xI0 (x)
qVL
where x = kT , and
q
gm = (IE0 + IEIF cos (ωIFt)) (4.55)
kT
and
VIF
IEIF = (4.56)
RE
q 2I1 (x)
iC (t) ' (IE0 + IEIF cos (ωIFt)) VL cos (ωLt) (4.57)
kT xI0 (x)
Finally, the output voltage of the mixer (if the RLC circuit is tuned to the sum frequency)
becomes
RLVIF I1 (x)
vout ' cos ((ωL + ωIF )t) (4.58)
RE I0 (x)
Mixers introduce a large amount of mixing products within the frequency spectrum
of the output current by virtue of device nonlinearity. However, the desired signal
is usually ωRF − ωLO component which is selected by the tuned band-pass filter.
This nonlinearity generates mixing products by two main sources. First, two adjacent
interferers may cause an undesired signal atop the desired signal due to IM3 component
as described before. Secondly, considerable leakage of LO and RF at the output causes
difficulties in extracting the desired signal. Another way to mix the two signals can be
implemented by applying both LO and RF signals to the base of a bipolar transistor.
Similarly, we can apply the LO signal at the emitter of a bipolar transistor and the
RF signal to its base. Finally, the exponential I −V characteristics of the device will
produce our desired mixing product. Figure 4.24 depicts a differential implementation
of a bipolar mixer which is single-balanced.
It should be noted that these mixers can also be implemented by MOS devices.
The important point in Figure 4.24 is the need for lower signal amplitude to achieve
the switching of Q2 and Q3 . In fact, in these devices, the RF current is applied to each
branch with a rate of LO signal. It can be roughly with a voltage of (VLO ) between
100 mV and 500 mV, the upper tree can be switched efficiently. In Figure 4.24, the RF
signal in each cycle appears at either of output terminals, while the other terminal is
grounded. Thus, we can assert the RF signal is multiplied by +1 or −1 alternatively.
Our objective is to obtain an equation for the output of the single-balanced mixer in
which the RF signal appears in common mode in the differential output. Since the
4.5 Mixer types Based on Switching Circuits 191
C R L L R C L R C
Vout
+
Vout
+ Q2 Q3
VLcosωLt
+1 or -1
Q1 Q1
+ +
VRcosωRt VRcosωRt
VBB1 RE VBB1 RE
LO drive is 180◦ out of phase, with respect to the base terminals, its leakage will be
present at the differential output. One may obtain the output of the single-balanced
mixer of Figure 4.24 as follows
and
Here ∗ sign stands for the convolution in the time domain or equivalently multiplication
of the corresponding impedances and current harmonics in the frequency domain. The
currents in each branch of the upper tree can be described as
IC VL cos (ωLt)
iC2,3 = 1 ± tanh (4.62)
2 2VT
VCC
RL CL RL
Q3 Q4 Q5 Q6
V2
Q1 Q2
V1
IEE
Example 4.1 Since the mixer circuit has three ports, how can we define power
in its ports? Moreover, discuss the IM3 component for two alternate interfering
channels with different spacings.
Solution:
Consider Figure 4.27.
4.5 Mixer types Based on Switching Circuits 193
Rin Rout
Rs RL
+
VRF
R1
+
VLO
If we have matching at the input and at the output, we have Rin = Rs , Rout = RL ,
and then we may write the conversion power gain
VIF 2
2RL
GP = (4.68)
VRF 2
2Rin
Suppose that the desired channel resides at 900 MHz and the LO signal is at
945 MHz. Therefore, the IF signal will be at 45 MHz.
(a) Consider the interfering channels are at f1 = 901 MHz and f2 = 902 MHz,
and (b) imagine the two interfering channels are at f1 = 900.03 MHz and
f2 = 900.06 MHz. Both of these channels could make IM3 components (e.g.,
ωLO − (2ω1 − ω2 )) atop of the desired signal. One way to mitigate this issue is the
implementation of a band-pass filter at the mixer’s input to eliminate those inter-
fering channels. Figure 4.27 depicts the structure in this case. Using a band-pass
filter with 1 MHz bandwidth, it is possible to eliminate the interfering channels
in the case (a). But having a band-pass filter of 60 kHz bandwidth at 900 MHz is
practically impossible and eliminating the interfering signals would become impos-
sible in case (b) at this stage (in this case, either the linearity of the mixer should be
improved or the wireless standard should require the levels of the adjacent channels
to be less than a predetermined value).
Rin Rout
Rs RL
+
VRF
R1
+
VLO
Applying two input signals as in Figure 4.29 and substituting in Equation 4.69, we
then obtain
1000Ω 1500Ω
50Ω 50Ω
R1
+ +
VRF Rs= VLO R L=
50Ω 50Ω
900MHz 45MHz
Figure 4.28: Typical matching circuit for a mixer, step-up capacitive input
matching, and step-down capacitive output matching.
4.7 Calculating IIP3 in Nonlinear Amplifier/Mixer 195
i(t) i(t)
+ +
V2cosω2t V2cosω2t
+ +
V1cosω1t V1cosω1t
VBB VBB
Figure 4.29: Applying two large signals to a nonlinear device (bipolar or MOS
transistors) to compute the compression point and the third-order intercept
point.
3 3 3 2
+ α1 A2 + α3 A2 + α3 A2 A1 cos(ω2t)
4 2
3α3
+ A2 A21 (cos((2ω1 + ω2 )t) + cos((2ω1 − ω2 )t))
4
3α3
+ A1 A22 (cos((2ω2 + ω1 )t) + cos((2ω2 − ω1 )t))
4
α3 α3
+ A31 cos(3ω1 )t + A32 cos(3ω2 )t
4 4
As stated earlier, one of the important parameters in the nonlinear amplifiers is their
measure of linearity which is obtained by means of a two-tone test. In this test, by
increasing the amplitude of tones, the output will be compressed and the low-level
slopes of the first- and the third-order terms will intersect at a point which we call
the intercept point. The term shown in Equation 4.71. is called signal amplitude
compression term which has a nonlinear relation with the input level and causes the
decrease in the amplifier gain as the input level is increased. Figure 4.30 illustrates
two different curves, one traces the fundamental harmonic term at the output as a
function of input level, and the other illustrates the output third-order intermodulation
amplitudes as a function of the input too, both on the log–log scale.
The −1 dB compression point is a point where the output fundamental level is
1 dB less than the presumed linear fundamental output level. Given A1 = A2 = A, the
compression point is computed as
3 3 −1
α1 A + α3 A3 + α3 A3 = 10 20 .α1 A (4.72)
4 2
196 Chapter 4. Mixers
Log(i)
20log(α1Ain)
OIP3 20log(0.75α3Ain3)
1dB
Log(Ain=A1=A2)
Pin,1dB IIP3
Figure 4.30: Output current of the amplifier versus its input signals’ amplitudes.
or
9 α3 2
A = −0.11 (4.73)
4 α1
Note that for the nonlinear amplifier to be compressive, we should have
α3
<0 (4.74)
α1
Otherwise, for αα31 > 0, the amplifier would be expansive which is generally a nonphys-
ical amplifier. Therefore, in the case compressive (physical) amplifier, we would have
r
α1
A1 dB = 0.22 − (4.75)
α3
Note that this is the “two-tone” compression point. It can be easily shown, by putting
A2 = 0, that a single-tone compression point can be expressed as
r
α1
A1 dB = 0.38 − (4.76)
α3
Verification of the above equation is left to the reader. From Figure 4.30, we are able
to compute IIP3 by the intersection of the two linear terms (tangents) as
3α3 3
20 log (α1 Ain ) = 20 log − Ain (4.77)
4
which finally gives the corresponding amplitude as
s r
4 α1 4 α1
AIIP3 = = − (4.78)
3 α3 3 α3
4.7 Calculating IIP3 in Nonlinear Amplifier/Mixer 197
i(t) = α0 + α1 v + α2 v2 + α3 v3 + α4 v4 (4.79)
Therefore, the desired mixer output will have the following form
3 α4 2 3 α4 2
vout = α2V1VS RL 1 + V + V cos ((ω0 − ωS )t) (4.83)
2 α2 S 2 α2 1
As it is obvious from the above equation, the output signal is compressed both with
respect to VS and with respect to V1 . So we define the −1 dB compression point as a
two-variable equation as follows
3 α4 2 3 α4 2 −1
1+ V + V = 10 20 = 0.891 (4.84)
2 α2 S 2 α2 1
Or
2 α2
VS2 +V12 = −0.11 × (4.85)
3 α4
198 Chapter 4. Mixers
9
Į
U
Į
U
96
Figure 4.31: The locus of the saturation voltages in the VS -V1 plane.
Or in another form
r
1 α2
VS2 +V12 2 = 0.269 − (4.86)
α4
This describes the compression effect in a mixer, which depends both on the LO
level and the signal level. This equation also describes a circle in the V1 ,VS plane
(Fig. 4.31). For example, if one considers the signal as a small input, he/she would
obtain the compression point by putting VS = 0 in the above equation, and obtain
the compression point in terms of V1 . Otherwise, if one considers the LO as a small
signal, he/she would obtain the compression point by putting V1 = 0 in the above
equation, and obtain the compression point in terms of VS . In a more general manner,
one can consider any proportion between V1 and VS , and compute the compression
point through Equation 4.86.
By sorting out only the desired output components at ω0 − ωS1 and at ω0 − ωS2 , we
will have
3 α4 2 α4 2 3 α4 2
vout =α2V1VS1 RL 1 + V + 3 VS2 + V cos ((ω0 − ωS1 )t)
2 α2 S1 α2 2 α2 1
(4.89)
3 α4 2 α4 2 3 α4 2
+ α2V1VS2 RL 1 + V + 3 VS1 + V cos ((ω0 − ωS2 )t)
2 α2 S2 α2 2 α2 1
4.7 Calculating IIP3 in Nonlinear Amplifier/Mixer 199
2 2 2 α2
VS1 + 2VS2 +V12 = −0.11 × (4.90)
3 α4
and
2 2 2 α2
VS2 + 2VS1 +V12 = −0.11 × (4.91)
3 α4
These two equations describe the compression phenomenon in a nonlinear mixer for
two-tone excitation. If one considers VS1 = VS2 = VS , the above equations simplify to
the following
2 α2
3VS2 +V12 = −0.11 × (4.92)
3 α4
Or in another form
r
12 α2
3VS2 +V12 = 0.269 − (4.93)
α4
This describes an elliptical contour in the V1 − VS plane. That is the contour which
describes a predetermined compression value (here, 1 dB) in the V1 −VS plane.
Normalized differential
VCC
collector current
∆I/IEE
ZL ZL
1
I1 I2
Vout 0.5
+
Q1 Q2
+ v/Vt
v -10 -8 -6 -4 -2 2 4 6 8 10 Normalized
-0.5 differential
IEE -1 input voltage
Figure 4.32: A typical differential bipolar stage and its corresponding nonlinear
transfer characteristics.
I
I1 = EE (4.98a)
1 + exp − V cos(ωt)
Vt
I
I2 = EE (4.98b)
V cos(ωt)
1 + exp Vt
and subtracting the first equation from the second equation in Equation 4.98, we obtain
the differential current as
V cos(ωt)
∆I = IEE tanh (4.99)
2Vt
If we employ the Taylor expansion of tanh {.} as
x3
tanh(x) = x − +··· (4.100)
3
Now, the current–voltage characteristic becomes
3
V 1 V
∆I = cos (ωt) − cos (ωt) + · · · (4.101)
2Vt 3 2Vt
which results in
v
u
u4 1
2Vt
AIIP3 = t 1
= 4Vt ≈ 100 mV (4.102)
3
24Vt 3
4.7 Calculating IIP3 in Nonlinear Amplifier/Mixer 201
Normalized differential
VDD drain current
∆I/I0
ZL ZL 1
I1 I2
Vout 0.5
+
M1 M2 v/(VGS0-VTH)
+
v -5 -4 -3 -2 -1 1 2 3 4 5 Normalized
-0.5 differential
I0 -1 input voltage
Figure 4.33: A typical differential MOS stage and its corresponding nonlinear
transfer characteristics.
As another example for a MOS differential pair, one can write with a good
approximation
v 2 v
I1 = k VGS0 + −VTH for < VGS0 −VTH (4.103)
2 2
and
2
−v v
I2 = k VGS0 + −VTH for < VGS0 −VTH (4.104)
2 2
Note for 2v > VGS0 −VTH , one of the transistors goes to saturation and the other one
goes to cut-off. Here VGS0 is the common DC bias voltage of either of the transistors
whose value is obtained by the following
r
I0
VGS0 = VTH + (4.105)
2k
Then
2
I1 VGS0 + 2v −VTH
= 2 (4.106)
I2 VGS0 − 2v −VTH
Given
I1 + I2 = I0 (4.107)
Then
" 2 #
VGS0 − v/2 −VTH
I1 1+ = I0 (4.108)
VGS0 + v/2 −VTH
202 Chapter 4. Mixers
and
I0
I1 = 2 (4.109)
VGS0 −v/2 −VTH
1+ VGS0 +v/2 −VTH
similarly
I0
I2 = 2 (4.110)
VGS0 +v/2 −VTH
1+ VGS0 −v/2 −VTH
Finally, the differential output current would have the following form, for 2v <
VGS0 −VTH
v
VGS0 −VTH
∆I = I1 − I2 = I0 (4.111)
v2
1+
4(VGS0 −VTH )2
Otherwise
I1 = I0 v
for > VGS0 −VTH (4.112)
I2 = 0 2
Vout,Diff
BJT
MOS
Vin,Diff
4Vt
2.3(VGS0-VTH)
R R R R
I I
Differential
collector current
∆I
IEE
1/RE
-REIEE
vin
REIEE Differential
input voltage
-IEE
Then
v
∆idd = id1 − id2 = for |v| ≤ RS I0 (4.124)
RS
∆idd = id1 − id2 = I0 for |v| > RS I0 (4.125)
The overall transfer characteristics of the differential MOS stage are depicted in
Figure 4.37.
A drawback of the structure depicted in Figure 4.35 is reduction in voltage head-
room which is wasted on these resistors. Therefore, we can modify Figure 4.35 to
solve this problem which is shown in Figure 4.38.
This structure doesn’t consume DC power in the resistor and it is a good prototype
for a more linear mixer. Note that in any case the conversion gain of the mixer with
degenerative resistors would be reduced with respect to nondegenerative mixer. That
is the conversion gain of the mixer goes from a maximum value obtained for RE = 0
or RS = 0 to zero for RE 1/gm or RS 1/gm .
Differential
drain current
∆I
I0
1/RS
-RSI0
vin
RSI0 Differential
input voltage
-I0
2R 2R
Example 4.2 Find the input and output power alongside conversion voltage
gain for the single-balanced downconverting mixer. Here the capacitors, C, are
considered to be short-circuit at RF and the LO frequencies, and they are considered
as open-circuit at the IF frequency. Furthermore, RC is small compared to RL .
9&& 9&&
5& 5/ 5/ 5&
& &
9RXW
4 4
9/2FRVȦ/2W
56
4
95FRVȦ5W
9LQ
9%%
Solution:
For computing the input power, we can write
1
Vin = VR (4.126)
1 + gin RS
Vin2rms
Pin = ginVin2rms = (4.127)
Rin
Given the fact that C in open at the IF frequency, for the output power, we can write
2
V√
out
2
Vout
2 2
PoutIF = ×2 = (4.128)
RL 4RL
Finally, for the conversion gain, given the fact that the upper tree is switched at the
LO rate,one may write
4 1 2
AV = gm RL = gm RL (4.129)
π 2 π
4.8 Linearization Methods in Mixers 207
Example 4.3 In the given mixer circuit, the LO is at 2.4 GHz with 1 V differential
for the upper tree and the RF frequency is at 2.41 GHz.
(a) With 1 mV signal for RF amplitude, find the IF component amplitude.
(b) If the input amplitude for RF signal is 1 mV, find the capacitor C in order to
attenuate the adjacent channel with the same amplitude residing at 2.45 GHz by
6 dB. The desired channel bandwidth is 2 MHz.
(c) Calculate the amplitude of LO signal without the capacitor C at the output.
(d) Suppose the double-balanced Gilbert cell and write its advantage.
(e) While in single-balanced given circuit only one branch has the RF current, why
the RF leakage is zero?
4 4
9/2
,(
4 ȍ
P$ & 95)
Solution:
(a) Considering complete switching of the differential pair, we have
2
gC = gm (4.130)
π
KT
rin = = 52 Ω (4.131)
qIE
rin 1 2 1 mV 2 0.5 mA
Vin = VBE ≈ VRF VIF = gm R × = 3 k×0.5 mV
rin + RS 2 π 2 π 25 mV
= 19.1 mV (4.132)
where ωc is the cut-off frequency of the output filter. Consequently, the value for
capacitor will be
208 Chapter 4. Mixers
2 2
C= = = 2.12 pF (4.134)
Rω2 3000 (2π) 50 106
(c) For the LO leakage (without consideration of load capacitance, the LO will
have a square-wave form, then), we have
Vout = R IEDC + IRF cos ω0t S (ω0t) (4.135)
4 4
Vout = R IEDC + IRF cos ω0t cos ω0t − cos 3ω0t + · · · (4.136)
π 3π
where
VRF
IRF = (4.137)
50 + re
Then
4
VLO = RIEDC = 1.9 V (4.138)
π
(d) An important feature of double-balanced Gilbert cell is removing the LO and
RF leakage to the output.
(e) As it is seen from the above equations, only the LO and the mixing components
appear at the output and because the RF signal is common mode, no RF signal will
emerge at the output (even without C).
One of the most practically applied mixers is the MOS Gilbert cell shown in Figure 4.41.
This mixer doesn’t show second-order nonlinearity due to its symmetry. Moreover,
VDD
RL RL
M3 M4 M5 M6
+
V2
-
M1 M2
+
VRF VRF
2 2
RS
VBB
since it has no current source at the source of input devices, it provides a larger linear
range operation for the RF signal.
Replacing x(t) by a two-tone input signal, and computing y1 (t) as before, and then
replacing y1 (t) by the computed result, we would obtain a sinusoidal expansion for
y2 (t). Then, by following the same procedure as described in section 4.7, it can be
shown that IIP3 voltage will be
s
4 α1 β1
AIP3 = 3
(4.140)
3 α3 β1 + 2α1 α2 β2 + α1 β3
1 1 α1 2 3α2 β2
≈ + + (4.141)
A2 IP3 A2 IP3,1 A2 IP3,2 2β1
Note that, given the fact that most of practical mixers/amplifiers use differential pairs
which have odd symmetry in their transfer characteristic, chances are that α2 and β2
are nearly zero. So, the third term in Equation 4.141 could be neglected with respect
to the first two terms. Equation 4.141 gives us an explicit equation to obtain IIP3 of
two cascaded stages. An important point is the effect of nonlinearity in subsequent
stages which will be more severe. We can compare Equation 4.141 with equivalent
resistance of parallel resistors, and by the assumption that the third term is neglected,
we can generalize Equation 4.141 to give the equivalent IP3 point for multiple stages
(here for three stages or more) as Equation 4.142:
1 1 α1 2 α1 2 β1 2
≈ + + +··· (4.142)
A2 IP3 A2 IP3,1 A2 IP3,2 A2 IP3,3
Here, it could be seen that the total IIP3 of cascaded stages is lower than each of them
in Equation 4.142. In other words, by the assumption of sufficient gain for previous
stages, the total IIP3 will be lower than the third stage IIP3 divided by previous gains.
210 Chapter 4. Mixers
In the above equation AIIP3 being the signal (voltage) amplitude and α1 and β1 being
the voltage gains, for a fixed input impedance system (e.g., 50 Ω), one can reexpress
Equation 4.142 in another form in terms of IIP3 powers and the power gains of the
stages as follows
−1 −1 −1 −1
PIIP 3 ,total
= PIIP3 ,1
+ PIIP G + PIIP
3 ,2 P,1
G G +...
3 ,3 P,1 P,2
(4.143)
Example 4.4 The given architecture is for a global positioning system receiver.
The received signal frequency is 1575 MHz and has a 2 MHz bandwidth and its
power is −130 dBm. The LO frequency is 1579 MHZ which downconverts the RF
signal to a 4 MHz carrier. With the given specifications, find
(a) The overall noise figure and overall IIP3 .
(b) If the mixer is linear, with the given two interferer signals, how much the
interferers’ IM3 component is lower than the desired GPS signal.
(c) We consider the effect of IIP3 of the mixers and the LNA, what is the input
interferer signal level which results in an IM3 component with −140 dBm power
level at the output.
The system impedance is 50 ohms.
*Y G%
1) G%
,,3 G%P I 0+]
0+] ,
%: 0+]
$
I 0+]
0+] 4
G% 1) G%
,,3 G%P ʌ %: 0+]
* G%
G%P
/2
G%P
I0+]
Figure 4.42: Typical GPS receiver architecture and the neighboring interfer-
ing signals.
4.9 Calculating Third-Order Input Intercept Point in Cascaded Stages 211
Solution:
(a) For the noise figure of cascaded stages one can write (see cascaded noise figure
expression in section 9.5)
100.8 − 1
F2 − 1
F = L × F1 + = 100.2 × 100.3 +
G1 101.5
= 1.58 × (1.99 + 0.16) = 3.42 (4.144)
FdB = 10 log (3.42) = 5.35 dB
AIIP3 2
2R
IIP3 = 10 log (4.145)
1 mW
1 1 G1 2
= + (4.147)
AIIP3tot 2 AIIP31 2 AIIP32 2
Given the fact that there is a 2 dB loss ahead of the LNA, then the IIP3 of the
combined filter and LNA would be
26
1 1 G1 2 1 10 20
= + = −10 −1 + −20 −1 ⇒ AIIP3tot = 7.06 mV
AIIP3tot 2 IIP3
1 −1
IIP3
2 −1 10 10 10 10
10 10 10 10
(4.149)
A2IIP
3tot
2R
IIP3tot = 10 log = −33 dBm (4.150)
1 mW
212 Chapter 4. Mixers
(b) At the LNA input, we have a pair of −57 dBm interfering signals. From
Equation 4.77, it can be seen that
3
AIM3 = α3 A3 (4.151)
4
or it can be rewritten as
3 α3 1
AIM3 = α1 A3 = α1 A3 (4.152)
4 α1 AIIP3 ,LNA 2
Now, the amplitude of the IM3 component, at the LNA output, can be computed as
q 3
1 3 1 15 −57 −1
AIM3 = α1 A = −12 −1 10 20 × 10 10 = 79.4 nV (4.153)
AIIP3 ,LNA 2 10 10
The desired GPS signal level at the output of the LNA (with 13 dB total gain)
would be
q
−117
AGPS = 10 10 −1 = 447 nV (4.154)
Since
447
20 log = 15 dB (4.155)
79.4
Therefore, the desired signal is 15 dB higher than the third-order IM of the interfer-
ing signal.
(c) To obtain an IM level of −140 dBm at the output, we should have
3 α3 1 1 33
AIM3 = α1 Ain 3 = 2
α1 Ain 3 = −33 10 20 × Ain 3 = 31.6 nV ⇒
4 α1 AIIP3 ,tot 10 10 −1
(4.156)
Ain = 32.9 µV
Ain 2
!
2R
Pin = 10 log = −79.7 dBm
1 mW
computation time. The important point is to decrease the ratio of the highest operating
frequency of the signal to its lowest frequency component. Once we have two exciting
tones with small frequency difference, their beat frequency will be very small. As such
in simulations related to IIP3 , it is recommended to increase the frequency distance
between the exciting tones. By this, we will avoid the requirement of too many points,
in the simulation, to differentiate between the frequencies of the exciting tones (and
consequently the beat frequency).
4.11 Conclusion
In this chapter, we have studied the different mixer topologies normally used in RF
circuits. The main application of the mixer block is to downconvert RF signal to IF for
detection in the receiver, or upconvert the IF signal to the RF in the transmitter.
It was shown that each mixer operates by virtue of its nonlinearity or switching
characteristics at the LO rate; however, evaluating the generated components needs
careful considerations to suppress unwanted mixing products. Moreover, a parameter
was introduced which is a measure of nonlinearity and was named 1 dB compression
point. Furthermore, in the presence of multiple input signals, another quantity was
introduced as the IIP3 , for computation of which analytical relations were presented.
Three different mixer topologies were studied, namely, unbalanced, single balanced,
and double balanced, and relations for port-to-port signal conversion were carried out.
Finally, methods to improve linearity in mixers by means of degeneration resistors
were investigated.
4.13 Problems
Problem 4.1 In the mixer circuit depicted in Figure 4.43,
1. Determine the output IF signal amplitude at 10.7 MHz for the case where the
RF current is iRF = 100 µA sin (2π × 100 MHz × t) and the LO signal is
VLO = 300 mV sin (2π × 89.3 MHz × t).
2. If the input signal has two components of the same amplitude one residing at
100 MHz and the parasitic one at 111.3 MHz, find the required 3 dB bandwidth
of the output low-pass filter. In order that the downconverted component of
the parasitic signal is 6 dB lower than the desired IF component, in this case,
calculate the appropriate value of C.
3. If the output low-pass filter has a 3 dB bandwidth of 11 MHz, find the parasitic
LO component at the output.
4. If the input has two components of 100 MHz and 100.1 MHz, compare the
conversion gain in dB for the output components at 10.7 MHz and 10.8 MHz
in comparison with the gain of IM3 components residing at 10.6 MHz and
10.9 MHz. In this case, we have
VRF = 50 mV sin (2π × 100 MHz × t)+50 mV sin (2π × 100.1 MHz × t) and the
bias current is 500 µA.
3V
RL=1kΩ RL=1kΩ
C
+ Q2 Vout Q3
vLO
-
iRF
Q1
+
VRF
ibias=500µA
VBB
Problem 4.2 In the circuit in Figure 4.44, a single-ended mixer with the input and
output matching networks is depicted.
1. Determine the values of L1 and C1 in order to match the RF input to 50 Ω, assume
that the input impedance of the transistor is 2 kΩ. Furthermore, determine the
values of L2 and C2 in order to match the output to 50 Ω. Suppose that the
output impedance of the transistor is 200 Ω. The RF frequency is 1900 MHz
and the IF frequency is 200 MHz. Assume that the capacitances CB1 and CB2
are RF short-circuit.
2. If the signals at the base of the transistor are VRF < VT and VLO > 10VT , calculate
the IF output current alongside the RF leakage current at the output in terms of
4.13 Problems 215
transistor’s gm and find an expression for the IF output voltage and the output
leakage voltage in this case.
Note that the RF trap circuit is open circuit at the RF frequency and short circuit at
other frequencies, and the the LO trap circuit is open circuit at the LO frequency and
short circuit at other frequencies.
VCC
R2 C
R1 B2
CB1
L1 L2 C2
IF-out
Q1
50Ω C1 50Ω
LO
RF
vRF trap
trap
50Ω
vLO
Problem 4.3 In the given mixer circuit depicted in Figure 4.45,assume the I − V
characteristics are described by i1 − i2 = 0.4vRF − 0.01vRF 3 tanh v2VLOT ,
1. Considering two signals of 50 mV amplitude at 104 MHz and 104.1 MHz at the
RF input, obtain the output components at 10.7 MHz and 10.6 MHz, consider-
ing a rectangular LO voltage in the upper tree (how?). The LO frequency is
114.7 MHz.
2. Compute the IM3 components at the output, and obtain the IIP3 point.
VCC
i1 RL=5kΩ RL=5kΩ i2
Vout
Q3 Q4 Q5 Q6
VLO
Q1 Q2
VRF
IEE
Problem 4.4 In the circuit depicted in Figure 4.46, determine the IIP3 through
ADS computer simulation. To save the computation time, use two distant tones with
150 MHz and 155 MHz frequencies as an example. The LO frequency is 225 MHz.
Note that the output low-pass filter has a cut-off frequency of 75 MHz which affects
the outputs. If one employs two close tones with 150 MHz and 150.1 MHz frequencies,
for example, he/she might obtain the same results with a much larger computation time
(why?). Choose bipolar transistors with a fT greater than 5 GHz in this simulation.
+9V
2.12pF 1kΩ 1kΩ 2.12pF
47Ω 47Ω
Q3 Q4 Q5 Q6
LO
220pF 220pF
VRF+ Q1 50Ω 50Ω Q2 VRF-
50kΩ 50kΩ
1mA
-9V
Problem 4.5 Consider the differential pair MOS mixer circuit depicted in Figure 4.47
where the RF input signal is applied to the gate of M1 , the LO signal is applied between
the gates of the differential pair, and the output-tuned circuits are tuned to the difference
frequency. Considering the RF signal as VS cos (ωSt) and the LO signal as V0 cos (ω0t).
Find an expression for the output IF signal. Here consider that the RF signal is a small
signal and the LO signal is a large signal with V0 ≤ 14 (VGS0 −VTH ) where VGS0 is the
bias voltage of the MOS differential pair transistors. All the MOS transistors are biased
in the active region.
9''
/ & & /
7XQHGWRȦȦ6 7XQHGWRȦȦ6
4 4
9RXW
0 0
Y
&
0
Y6 5 5
9''
Problem 4.6 In the downconverting mixer depicted in Figure 4.48, consider β of the
transistors is sufficiently large such that 2rπ ≥ 25 kΩ, and L = 765 nH with QL = 10.
Find the appropriate values of C1 and C2 for matching the 50 Ω RF source to the input
of the mixer. Furthermore, determine the required value of I to achieve a conversion
(10.7 MHz)
power gain GPC = PPout(104 MHz) = 12 dB. Note that the RF signal at the input of the
in
center tapped capacitor is vRF = 1 mV cos (2π × 104 MHz × t). Moreover, the quality
factor of the capacitors is assumed to be large.
VCC
Q3 Q4 Q5 Q6
LO
C∞
Q1 Q2
50Ω C2 C∞
L 5kΩ I 5kΩ
C1
1.5V 1.5V
mV MHz
vLO=200 cos(2π×114.7 t)
Figure 4.48: Gilbert cell double-balanced mixer with input matching.
Problem 4.7 In this problem, we try to learn how to simulate the IIP3 point deter-
mination. In order to have a lower simulation time, we should choose two-tones far
enough for more time-efficient simulation. To begin the simulation (e.g., by ADS or
Cadence IC design), run the transient simulation for two different low-power signals
to obtain points A and B depicted in Figure 4.51. At the same time, find points C and
D through the IM3 component computation. By extrapolating the two straight lines,
you may find the IIP3 point. Note that a common-mode 2.5 V bias is applied to the LO
port. Furthermore, that you can use the following relations in your simulations:
2 2 2
(Va ) (Vb ) (Vout )
va = Va cos 2π (103 MHz)t, vb = Vb cos 2π (102 MHz)t, Pin = 8(R = 8(R , Pout = 2(R
in ) in ) out )
where Rin = 50 Ω and Rout = 1 kΩ.
Hint: choose a small value for Va and Vb while Va = Vb , and increase both of them
gradually.
218 Chapter 4. Mixers
5V
1kΩ 1kΩ
+ V -
out
Q3 Q4 Q5 Q6
LO
1.5V 1.5V
LO=200(mV)cos(2π×114.7(MHz)t) LO-CM=2.5V
Figure 4.49: A matched Gilbert cell double-balanced mixer to test the IIP3 .
RFC
+ LO+
C∞ +
VLO/2
VLO -
VDC +- RFC - +
VDC +- VLO/2
-
C∞ LO-
(a) (b)
G%'
VORSH
3RXWG%P G%'
VORSH
%
$
'
&
3LQG%P
,,3
Problem 4.8 If in an amplifier for the input signal pair of −70 dBm level, we obtain
output signals of −50 dBm and output IM3 components of −80 dBm,
1. Determine the IIP3 of a single-stage amplifier in dBm and its gain in dB (Hint:
use formula IIP3 = ∆p2 + pin ).
2. If two similar stages of the same amplifier are cascaded, determine the overall
IIP3 in dBm, and the output IM3 components in dBm in case of −70 dBm input
signals.
Problem 4.9 If the output spectrum of the first mixer in the receiver chain depicted
in Figure 4.52 is like what is shown in the figure,
1. Determine the IIP3 for the first mixer, if it has a conversion gain of 10 dB. (Input
signals reside at 899.97 MHz and 899.94 MHz. Moreover, fLO,1 = 945 MHz,
and fLO,2 = 45.455 MHz).
2. Assuming the second mixer having the same nonlinear characteristics as the
first mixer, find the spectrum of the output of the third filter and then given
unequal input signals, find the IM3 components at the output of the second
mixer (the transfer function of the third filter is also shown).
3. Find the output spectrum of the fourth filter.
A B C D E F G
1st filter 2nd filter
fLO1 fLO2
|HBPF3| |HBPF4|
0dB 0dB
A B
-60dBm
-12dB/oct -12dB/oct
-90dBm
f(MHz) f(MHz) f(kHz)
45
45.03
45.06
45.09
Problem 4.10 In the given receiver depicted in Figure 4.53, determine the overall
noise figure and the overall IIP3 (see section 9.5 for the expression for the noise figure
of the cascaded stages). You may determine the noise figure and the IIP3 of the first
four blocks, then those of the last two blocks, and consequently the overall noise figure
and the overall IIP3 . Then, suppose that the desired signal power is −60 dBm at the
input. If two interferer signals both with a power of −50 dBm at 60 kHz off the desired
signal and 120 kHz off the desired signal, respectively, appear at the receiver input,
calculate their effect at the output of the IF amplifier. What kinds of unwanted signals
appear at the output of the IF amplifier? How much the unwanted signals are lower
than the desired signal?
220 Chapter 4. Mixers
_+%3)_
G%RFW
G%
ORJI
N+]
N+]
N+]
N+]
Figure 4.53: Receiver chain for determination of the overall noise figure and
overall IIP3 .
10V
1kΩ 5nF
Vout
Si
Q1 α≈1 Q2
+
vs
3.3kΩ
+ Si
v1 α≈1
-10V
Problem 4.12 Determine the main mixing component at the output. Furthermore,
calculate the LO leakage signal at the output of the mixer circuit depicted in Figure 4.55.
vs
+
Vout
+ +
v1 V 100Ω i C R L
i=0.1V+0.3V2+0.01V3 A L=50µH
V1=2cos(5×106t) V C=20nF
Vs=5[1+0.5f(t)]cos(9×106t) mV R=2kΩ
Problem 4.13 In the mixer circuit depicted in Figure 4.56, the LO signal is a square-
wave pulse train as depicted in the figure. Determine the output voltage at the sum
frequency and the unwanted component at the difference frequency.
v1(t) Rs
Vout
1kΩ
C R L
1V + +
t vs v1
T0=2π×10-8s
T0 vs=50(mv)[1+0.6cos(103t)]×cos107t L=826.4nH
MOS Parameters: C=100pF
VTH=1V R=2kΩ
K=2mA/V2
i = αV + βV 2 + γV 3 (4.157a)
where
α = 2 mA/V (4.157b)
2
β = 0.5 mA/V (4.157c)
3
γ = −0.2 mA/V (4.157d)
222 Chapter 4. Mixers
M12 V
L1 L2
C1 R1 L0 C0 R0
is
vLO
R1=2kΩ R0=1kΩ
C1=100PF C0=1nF
L1=1μH L0=2.5μH
L2=100nH vLO=3cos1.2×108t V
M12=300nH
Is=50(1+0.5f(t))cos108t μA
vs
Rs
+
Vout
+ +
v1 Ri v i C R L
Baseband signals are generally band-limited low-pass signals which cannot be directly
transmitted over the transmission medium. Furthermore, a huge number of different
signals should be transmitted simultaneously in a transmission medium (apparently
the air or other transmission medium), so it is imperative that the baseband signals to
be modulated over the radio carriers at different frequencies before being transmitted
over the air, with antennas of limited sizes. This allows different modulating signals
to be differentiated or to be distinct in the frequency domain. In this chapter, we
discuss the conventional modulation schemes alongside modern digital modulations
with high bandwidth efficiency. Moreover, we investigate the receiver structures for
the signal demodulation. One of the long-existing modulation schemes which is used
even today is the amplitude modulation (AM) for long-distance broadcasting. In this
method, the data are embedded on the amplitude of the signal, therefore it is sensitive
to amplitude noise. Furthermore, we introduce circuits to demodulate AM signals. One
of the applicable modulations is the phase modulation (PM). In this modulation, the
baseband data are embedded in the phase of the radio signal enabling high data rates.
We will then discuss quadrature amplitude modulation (QAM) which is one of the
most applied modulation schemes in modern radios, phase modulator and demodulator
in this chapter. Finally, a few special modulation schemes are investigated.
5.1 AM Modulation
We can represent a sinusoidal AM-modulated signal as follows
where A cos(ωCt) denotes the carrier signal, m is the modulation index, and cos(ωmt)
is the baseband signal. For AM modulation signal, modulation index is equal to or less
than unity. Figure 5.1 depicts a typical AM-modulated signal. As it is obvious from
Figure 5.1, the baseband signal is embedded as the envelope of the signal. Moreover,
224 Chapter 5. Modulation/Demodulation of Amplitude/Phase
1+m
1-m
Amplitude
spectrum
Baseband ωc
signal
f
LSB USB
signal amplitude is limited between 1 + m and 1 − m values. Notice that we can expand
Equation 5.1 to achieve
m m
VAM = A cos(ωCt) + A cos(ωCt − ωmt) + A cos(ωCt + ωmt) (5.2)
2 2
Equation 5.2 shows that three frequency components appear at the output which are
ωC ± ωm , and ωC . If we assume that the baseband signal has a finite spectrum (low
pass spectrum), the output will be similar to what is shown in Figure 5.2.
Figure 5.2 illustrates the fact that the baseband signal is upconverted around the
carrier signal, and also subscripts LSB and USB refer to the lower sideband and the
upper sideband, respectively. This shows that the baseband signal is present at both
sides of the carrier.
5.2 AM Demodulation
The easiest method to demodulate the AM signal is to extract the baseband signal from
the envelope of the received RF signal. This can be done using a diode and a capacitor
which is shown in Figure 5.3.
In Figure 5.3, the antenna receives the AM-modulated signal, and develops an AM
voltage at the diode input. The R −C low-pass filter extracts the low-pass components
of the rectified signal which is the envelope of the RF signal. Figure 5.4 illustrates the
behavior of the circuit presented in Figure 5.3. As in Figure 5.4, the envelope of the
5.2 AM Demodulation 225
R C
Figure 5.4: The concept of AM demodulation using a diode with R–C circuit.
1
Output signals (V)
-1
RC=50ns
RC=500ns
-2
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time (µs)
RF signal is proportional to the baseband data. As such, the nonlinearity might affect
the demodulation process. Figure 5.5 illustrates an AM signal with the modulation
index of unity.
In Figure 5.5, the carrier frequency is selected to be 10 MHz and the baseband
signal is assumed to be 1 MHz. A problem arises once the R-C time constant of
the output circuit is too long with respect to the period of the modulating signal. In
this case, a failure-to-follow distortion is caused once the R-C low-pass bandwidth is
insufficient. That is
1
≤ ωm (5.3)
RC
226 Chapter 5. Modulation/Demodulation of Amplitude/Phase
1
the output distortion will occur, so normally one should choose ωc RC ≥ ωm to have
less distortion.
qV1 V1
x= = (5.5)
kT Vt
VCC VCC
C R L L R C
Vout
+ -
iE1 iE2
+ Q2 Q3
V1cos(ωct)
-
Q1 IE
+
Vmcos(ωmt)
VBB RE
The collector currents of the transistor pair (Q2 and Q3) can be expressed as
IE x
iE1 = 1 + tanh cos (ωct) (5.6)
2 2
IE x
iE2 = 1 − tanh cos (ωct) (5.7)
2 2
The time-domain differential output voltage can be described as
where zL (t) is the impulse response of the load impedance, and ∗ stands for the
convolution process. The output voltage reduces to
or
x
Vout = −IE tanh cos (ωct) ∗ zL (t) (5.10)
2
The load impedance in the frequency domain can be expressed as
R
ZL ( jω) = (5.11)
1 + jQ ω
ωc − ωωc
For harmonic components of the input frequency, the load impedance can be described
in terms of harmonic frequencies (i.e., for the nth harmonic). Note that the output
tuned circuits should be tuned to ωc
R nR − jnR
ZL ( jnωc ) = = ≈ (5.12)
1 + jQ n − 1n n + jQ (n2 − 1) Q (n2 − 1)
Once the input RF voltage is sufficiently small (less than 50 mV for a bipolar differential
pair), the harmonics become negligible and the above expression is reduced to
IE (t)
Vout = − x cos (ωct) × R (5.14)
2
And the output in this case becomes the following which is apparently an AM-
modulated signal
R (VBB −VBEQ ) Vm qV1
Vout = − 1+ cos (ωmt) cos (ωct) (5.15)
2RE VBB −VBEQ kT
228 Chapter 5. Modulation/Demodulation of Amplitude/Phase
Amplitude
spectrum
Input baseband signal
ωm
Amplitude
spectrum Output spectrum with
the effect of BPF
ωc
3ωc
Figure 5.7: The input baseband spectrum and the output spectrum of an AM
modulator (the two sidebands and the carrier are distinct in the output spectrum).
5.4 Double- and Single-Sideband Suppressed Carrier Generation 229
where k is the mixer’s output proportionality factor. As Equation 5.20 suggests at the
output, we have two frequency components as ωc ± ωm and there is no effect of the
carrier signal itself. Since the AM signal has components at both sides of the carrier
(it has two sidebands) which transmit the same amount of information, the idea of
removing one of the sidebands comes to mind in order to have a more bandwidth-
efficient modulation. This modulation is called single-sideband suppressed carrier
(SSBSC). We can implement SSBSC by the block diagram shown in Figure 5.8.
Assuming sinusoidal baseband, for the sake of simplicity, the outputs of each of
the mixers become
π
V1 = kVcVm cos ωmt + cos (ωct) (5.21)
4
π
V2 = kVcVm cos ωmt − sin (ωct) (5.22)
4
and the total output after the summer becomes
h π π i
Vout = kVcVm cos ωmt + cos (ωct) + sin ωmt + sin (ωct) (5.23)
4 4
or
π
Vout = kVcVm cos (ωc − ωm )t − (5.24)
4
Vccosωct
-90º
V1
45º
Baseband
signal Vout
-45º SSB-AM-SC
V2
which is apparently an SSBSC signal. The structure in Figure 5.8 occupies half of the
bandwidth of the DSBSC AM signal; however, its drawback is the implementation
of precise wideband phase shifter circuits which should have more than two decades
of bandwidth, for example, from 100 Hz to 12 kHz. To mitigate this problem, we
can implement the phase shifters after the upconverters. Finally, Figure 5.9 can be
presented as a modified version of Figure 5.8.
As Equation 5.24 suggests, the carrier and the upper sideband are not present at
the output and as a result, a better spectral efficiency is achieved.
Figure 5.10 illustrates an AM modulator where the LO signal is not present at the
output (suppressed carrier) because of the symmetry of the circuit at the baseband.
Here, the bias current of the upper tree becomes a function of the carrier voltage:
VBB +Vc cos (ωct) −VBEQ
IE = (5.25)
RE
The output of the upper tree becomes
Vout = (iE2 − iE1 ) ∗ zL (t) (5.26)
Assuming Vm ≤ Vt , we can write
IE qVm
Vout = − cos (ωmt) × R (5.27)
2 kT
In another form
Vc qVm R Vc Vm R
Vout = − cos (ωmt) cos (ωct) = − cos (ωmt) cos (ωct) (5.28)
2RE kT 2RE Vt
The carrier signal is suppressed at the output because it is the common mode at the
upper tree as stated in Chapter 4.
Vccosωct
-90º
45º
Baseband
signal Vout
-45º SSB-AM-SC
Figure 5.9: Block diagram implementation of the SSBSC AM signal with phase
shifters at the carrier frequency.
5.5 Synchronous AM Detection 231
VCC VCC
C R L L R C
Vout
+ -
iE1 iE2
+ Q2 Q3
Vmcos(ωmt)
-
Q1 IE
+
Vccos(ωct)
VBB RE
CL RL RL CL
- Vout +
Q3 Q4 Q5 Q6
v2
Q1 Q2
v1
AM Signal IEE
Assuming small-signal inputs, such that the Gilbert cell functions in the linear range,
the output simplifies to
qv qv
1 2 v1 v2
Vout = IEE ∗ zL (t) = IEE ∗ zL (t) (5.31)
2kT 2kT 2Vt 2Vt
and
V1V2
Vout = IEE RL m cos (ωmt) (5.34)
8Vt2
If we assume a large signal for V2 (i.e., hard switching of the upper tree), the AM
signal is in effect multiplied by a square wave of the carrier frequency, and as such, the
carrier harmonics are multiplied by the AM signal and the low-pass component will
appear at the output with a form as follows
2 V1
Vout = IEE RL m cos (ωmt) (5.35)
π Vt
The output is clearly proportional to the modulating signal. Figure 5.12 illustrates the
demodulation process.
5.5 Synchronous AM Detection 233
Amplitude
spectrum
RF input spectrum
ωc
Amplitude
spectrum
IF spectrum after LPF
ωm
The AM signal voltage is limited to VL , by the limiting amplifier, so the input voltage
of the upper tree becomes
VDD
C RL RL C
+ Vout -
M3 M4 M5 M6
GA M1 Rs Rs M2
+ +
v1 GB v2
-
-
The low-pass component of the product of the above two voltages at the output becomes
k
Vout = V1VL (1 + m cos (ωmt)) (5.39)
2
Finally, the most applicable usage of Gilbert cell which is a mixer is achieved by
prefect switching of the upper tree and small-signal injection at the lower tree.
4
$HMij
,
π
11 : A cos ωct + (5.40a)
4
3π
10 : A cos ωct + (5.40b)
4
π
00 : A cos ωct − (5.40c)
4
3π
01 : A cos ωct − (5.40d)
4
Figure 5.16 depicts the RF symbols in QPSK modulation for a 4 MHz carrier (as an
example) and 1 M symbol per second (2 Mbit/s) data rate, where all of them have the
same amplitude; however, all adjacent symbols are orthogonal to each other.
Rb = SR log2 m (5.41)
Q Acos(ωct)
0 1
* * I
10 11
* *45
0
I
01 00
* *
Figure 5.16: Constellation of QPSK modulation.
5.7 Modern Practical Modulations 237
Acos(ωt+π/4)
1
0.5
Output signals (V)
11 10 00 01
-0.5
-1
0 0.5 1 1.5 2 2.5 3 3.5 4
Time (µs)
* * * *
* * * * I
* * * *
* * * *
Figure 5.18: 16 − QAM constellation.
where Rb is the bit rate, SR is the symbol rate, and m is the number of symbol levels
(normally, the transmission bandwidth, BW , is chosen about the symbol rate). This
modulation transmits four bits per symbol. Thus, 16-QAM has a higher bit rate with
respect to the QPSK within the same bandwidth. In this scheme, each four bits are
transmitted by a single symbol. For the symbols, there are 3 different amplitude levels
and 12 different phase levels as shown in Figure 5.18. In this case there are 16 distinct
symbols, and therefore, Rb = 4SR .
* * * * * * * *
* * * * * * * *
* * * * * * * *
* * * * * * * * I
* * * * * * * *
* * * * * * * *
* * * * * * * *
* * * * * * * *
Figure 5.19: 64 − QAM constellation.
VCC
C
Q2 Q3
RL
Baseband
Q1 IE
+
Vccos(ωct)
VBB RE
where f (t) is a random binary signal which varies between +1 and −1, and Vm is the
logic amplitude. The difference current of the differential pair (if VVmt < 1) becomes
IE Vm f (t)
∆iEE = (5.44)
2 2Vt
VcVm RL
Vout = f (t) cos (ωct) (5.45)
4Vt RE
which is apparently a BPSK signal. As such, in Figure 5.20, the baseband signal, which
is equal to ±1, is upconverted to RF frequency and goes through the air by the antenna.
Then, the coupled signal with the BALUN (balanced to unbalanced) feeds two antenna
branches.
5.7.6 Generating and Detecting the Quadrature Phase Shift Keying Signal
It is possible to have a structure to transmit the baseband in QPSK form. The required
architecture is shown in Figure 5.21. Each point on the constellation can be conceived
→
−
by a vector such that V = Aejϕ . As a result, Figure 5.21 makes it possible to generate
each point
√ of the constellation.
√ For instance, for QPSK modulation, we can choose
I = ± 2/2 and Q = ∓ 2/2 and the modulated signal would have the following form
√
2
vc (t) = ai (t) cos ωct − aq (t) sin ωct (5.46)
2
FRVȦFW
%3)
,
6HULDOWR
SDUDOOHO
%LQDU\ FRQYHUWHU 2XWSXW
GDWD 4
%3)
VLQȦFW
FRVȦFW
/3)
, %LQDU\
,QSXW 3DUDOOHOWR
GDWD
VHULDO
FRQYHUWHU
4
/3)
VLQȦFW
Likewise, an architecture as depicted in Figure 5.22 can be used for QPSK demodula-
tion. The output voltages of the demodulator mixers can be expressed as
√
2
vo1 (t) = KVc (t) × cos ωct = K ai (t) + ai (t) cos(2ωct) − aq (t) sin(2ωct)
4
(5.47a)
√
2
Vo1,LP = K ai (t) (5.47b)
4 √
2
vo2 (t) = K Vc (t) × sin ωct = K ai (t) sin(2ωct) + aq (t) − aq (t) cos(2ωct)
4
(5.47c)
√
2
Vo2,LP = K aq (t) (5.47d)
4
where K is the proportionality constant of the mixer.
In this architecture, the in-phase and the quadrature bit streams are synchronously
detected and they are applied to the parallel to serial converter at the output of the
low-pass filters.
%3)
, %3)
6HULDOWR
SDUDOOHO
'DWD FRQYHUWHU
4
%3)
VLQȦW
signal to be transmitted through the air. One of the main considerations in transmitter
design is quadrature mismatches which can be frequency- and time-dependent. These
errors cause the points on the constellation to change their phase/amplitude, and this
results in difficult signal detection. Moreover, it also degrades the detection probability
at the receiver. This phenomenon is shown in Figure 5.24. Here it is observed that
if the SNR is diminished, each symbol in the constellation might interfere with its
adjacent symbols, and therefore, introduce errors in the detection process.
Another assumption is the operation of mixers which act as ideal switches.
However, in reality, this may not happen and RF harmonics might be troublesome.
Figure 5.25 depicts the receiver for I/Q detection. Here it is assumed that the I and
the Q channels introduce an amplitude error of ∆G/2 and a phase error of ∆ϕ/2 on
each path.
Since the frequencies of transmission and reception in the direct-conversion sys-
tem are the same, the oscillators and mixers can be used for both purposes. This
architecture is called a coherent transceiver. Another drawback in this structure is
frequency variations due to temperature. Suppose the outgoing signal is at 900 MHz
which is generated by a temperature-compensated crystal oscillator (TCXO) which
has a 3 ppm frequency variation. This means that a frequency error of 2.7 kHz may
Q
High
SNR
* *
I
* *
Figure 5.24: Phase and amplitude error effect in QPSK modulation due to
noise, in two cases, low SNR and high SNR.
242 Chapter 5. Modulation/Demodulation of Amplitude/Phase
FRVȦWǻij
*ǻ* /3)
%3)
3DUDOOHOWR 'DWD
VHULDO
FRQYHUWHU
*ǻ* /3)
VLQȦWǻij
occur during the operation of the circuit. This frequency deviation will result in the
rotation of the points on the constellation which in fact increases the probability of
error in detection. To mitigate this issue, one of the solutions is using a PLL to lock the
phase and the frequency of the LO to the received signal. Finally, we can summarize
the signal distortions in the following
(1) Gain mismatch results in rectangular constellation distortion (shown in
Figure 5.26).
** ** Ideal constellation
I Rectangular constellation
** **
(∆ G/2 gain mismatch
in each path)
** ** Ideal constellation
I Parallelogram constellation
**
in each path)
SKDVHVKLIW
, ,
4 4
%LWSHULRG 7E 7E
7E
D436. E2436.
Figure 5.28: Signals in quadrature phase modulation, (a) a QPSK signal pair
with a 180◦ phase change, and (b) an OQPSK signal pair where the 180◦ phase
shift is avoided.
0DJQLWXGH 3KDVH
I I
*DXVVLDQOLNH /LQHDUSKDVH
FRVȦW
*)
%3)
6HULDOWR
SDUDOOHO
'DWD FRQYHUWHU
7E *)
VLQȦW
Magnitude QPSK
OQPSK
GMSK
Figure 5.31: The comparison of the spectra of QPSK, OQPSK, and GMSK
signals.
magnitude behavior and a linear phase response which is equivalent to a constant delay.
Figure 5.29 shows the frequency response of a Gaussian filter (GF).
A typical GMSK transmitter block diagram is shown in Figure 5.30.
Figure 5.31 shows a comparison of the spectra of the three types of quadrature
modulations. As it is obvious, the GMSK has the lowest side-lobes’ level.
Although, nowadays, analog amplitude and FM for legacy radio and television
broadcasting are used, there are modern digital receivers with compatible analog
techniques for their detection, as well. The following example describes this issue to
detect frequency-modulated signal with I/Q demodulator.
R
Example 5.1 If we consider FM-modulated signal as X = A cos ω0t+k Vm
0
0
cos ωmt dt +φ , the instantaneous frequency will be ω0 + kVm cos(ωmt). Sug-
gest an structure to demodulate an FM signal with quadrature zero IF receiver (note
that this can be done by two analog multipliers, two differentiators, and a voltage
subtractor). In cell phones, FM signals are demodulated with this structure with
digital signal processing right after the mixers.
Solution:
We can use the structure shown in Figure 5.32 for FM detection.
5.8 Effect of Phase and Amplitude Mismatch on the Signal Constellation 245
cos(ω0t)
LPF LPF
A D
d/dt -
Xout
d/dt +
B C
LPF LPF
sin(ω0t)
For the signals at nodes A and B, with the assumption of filtering out other
mixing products, we will have the low-pass components as
Z
A 0
0
XA = cos k Vm cos ωmt dt + φ (5.48a)
2
Z
A 0
0
XB = − sin k Vm cos ωmt dt + φ (5.48b)
2
Z
d A 0
0
XA = − kVm cos (ωmt) sin k Vm cos ωmt dt + φ (5.49a)
dt 2
Z
d A 0
0
XB = − kVm cos (ωmt) cos k Vm cos ωmt dt + φ (5.49b)
dt 2
Finally, signals in nodes C and D with the assumption of low-pass filter at the
output will be
A2
XC = kVm cos (ωmt) (5.50a)
4
A2
XD = − kVm cos (ωmt) (5.50b)
4
A2
⇒ Xout = XC − XD = kVm cos (ωmt) (5.50c)
2
Using digital signal processors, this process can be implemented in digital domain
as well, as such a digital receiver could be compatible with an analog modulation
technique.
246 Chapter 5. Modulation/Demodulation of Amplitude/Phase
Example 5.2 For the zero IF receiver shown in Figure 5.33, given the fact that
the in-phase and quadrature detectors’ carriers are locked to the input carrier, prove
that the output will be the AM detected signal if m < 1.
Solution:
Consider the given receiver in Figure 5.33.
FRVȦFW
/3)
;, ;,/3
%3)
9ROWDJH
;LQ ;RXW
DGGHU
;4 ;4/3
/3)
VLQȦFW
A A
Xout = cos (φ ) (1 + m cos (ωmt)) − sin (φ ) (1 + m cos (ωmt)) (5.53)
2 2
A
= (1 + m cos (ωmt)) (cos (φ ) − sin (φ ))
2
√
2A π
= (1 + m cos (ωmt)) cos φ +
2 4
The above equation suggests that the output signal is dependent on the phase of the
input AM-modulated signal. For instance, if this phase is 45◦ , the output will be
zero. To solve this problem, indeed we should use a PLL to lock the LO signals
to the transmitted carrier signal. In other words, the received signal is injected
to a PLL, and then the phase values of the in-phase and the quadrature LO input
of the mixers are chosen in order to make the maximum amplitude which is a
5.8 Effect of Phase and Amplitude Mismatch on the Signal Constellation 247
phase difference of −45 ◦ between the LO and the carrier signal. That is to say, the
√
maximum value of 22A (1 + m cos (ωmt)) is achieved for φ = −45.
In case of digital signal processor implementation of the detection process, one
could implement the following relation
q A
Xout = XI2 + XQ2 = (1 + m cos (ωmt)) (5.54)
2
Apparently here, the use of a PLL is not needed.
Example 5.3 For the given synchronous AM detector presented in Figure 5.34,
the modulated signal has a form presented in Figure 5.35. Calculate the output
detected signal amplitude and the unwanted second harmonic component at 2 MHz.
9
& 5/ 5/ &
Q) Nȍ Nȍ Q)
4 4 4 4
Y Y
Y
5 5
Y 4 5( ȍ 4 Y
5 5
P$ P$
9
9&0 9
t1=1ms
800mV
400mV
t2=1μs
Solution:
One can express the differential current of the Gilbert cell multiplier as (note that
here there is a linearizing resistance, RE = 500 Ω, between the emitters of the lower
tree)
IEE v1 v2
∆IEE = tanh (5.55)
1 + gm R2E 2Vt 2Vt
The large-signal input voltage of the upper tree switches the value of hyperbolic
tangent between +1 and −1 , so the differential output current can be expressed as
IEE v1
∆IEE = S (ω0t) (5.56)
1 + gm R2E 2VT
2 V1 (1 + m cos (ωmt)) gm
Vout = RL (5.58)
π 10 1 + gm R2E
It can be seen from Figure 5.35, the AM signal frequency is 1 kHz and the carrier
frequency is 1 MHz. The modulation index is also 0.333. The input voltage has the
following form
The unwanted second harmonic (2 MHz) component at the output can be calculated
by considering the second harmonic current and the load impedance at 2 MHz. The
second harmonic component amplitude is the same as the DC current amplitude
(why?). The load impedance for the second harmonic component becomes
RL
ZL ( jω) = (5.61)
1+ jω ω
cut−off
5.10 References and Further Reading 249
where
1
ωcut−off = = 2π × 10 kHz (5.62)
RLC
RL
ZL ( j2π × 2 MHz) ≈ − j (5.63)
200
The unwanted 2 MHz output voltage will take the following form
π
Vout (2π × 2 MHz) = 0.2 1 + 0.333 cos 2π × 103t cos 4π × 106t −
(mV)
2
(5.64)
5.9 Conclusion
In this chapter, the AM modulation and demodulation techniques as well as double-
sideband suppressed carrier AM and single-sideband suppressed carrier generation
were studied. Different digital modulation techniques such as BPSK, QPSK, and
QAM were presented as well. The quadrature digital modulator architecture as well as
quadrature digital receiver were studied. The synchronous AM detection was presented
and the quadrature demodulator was used for AM and FM detection as shown in the
examples. Normally, the best modulation technique is the one that has the higher data
rate within the specified bandwidth. The maximum achievable data rate is specified by
the information theory formula Rb = BW × log2 (1 + S/N). This means that the more
we increase the discrete levels of digital modulation (in order to transmit more bits of
data within a symbol), the higher S/N level we need to be able to distinguish between
different symbol levels.
5.11 Problems
Problem 5.1 In the QPSK transmitter/receiver system shown in Figure 5.36 using
MATLAB software,
1. Determine the transmitted spectrum with the random bit sequence generator at
the input of the baseband data. Assume that the LO is locked by a PLL.
5E .ELWVHF
5%6* 5DQGRPELWVHTXHQFHJHQHUDWRU
$FRVȦW
$FRVȦW
,W 4 4
I 0+] I 0+] ,U
6SOLWWHU
63
5%6*
FRQYHUWHU
4U
4W ,QSXW
,W
$VLQȦW
$VLQȦW
4W
Problem 5.2 An AM signal is applied between the bases of Q1 and Q2 in the analog
multiplier as shown in Figure 5.37,
9
& Nȍ Nȍ &
9RXW
4 4 4 4
/2
4 Nȍ 4
5)
P$ P$
9
/2 P9FRVʌî.+]W
5) P9FRVVLQȦEWFRVʌî.+]W
1. Determine the capacitor C such that the second carrier harmonic at the output is
30 dB lower than the desired signal.
2. Determine the baseband component at the output. Assume that ωb = 2π ×1 kHz.
Problem 5.3 In the 16-QAM modulator depicted in Figure 5.38, two double-balanced
analog multipliers are employed like the one used in problem 2, terminated on 3 kΩ
loads.
1. Find the necessary amplitude for the input I and Q channels to achieve a
maximum output level of 100 mV. The LO signal has an amplitude of 200 mV
at 1 MHz frequency. Furthermore, determine the output spectrum about 1 MHz
for a bit rate of 4 kbit/s. (Note that in 16-QAM each symbol represents 4 bits of
information).
2. If the quadrature signal has a finite phase error of 5◦ , draw the generated
constellation. Moreover, if the resistor on the Q input path has a 10% error (that
is, it turns into 1.1 kΩ), how the constellation will be deformed.
9
Nȍ Nȍ Nȍ Nȍ
9RXW 9RXW
4 4 4 4 4
4
4
4
/2, /24
4 Nȍ 4 4
Nȍ 4
, 4
9
4
WROHYHO
7R,LQSXW
FRQYHUWHU
63
'DWDLQ
FRQYHUWHU
,
WROHYHO
7R4LQSXW
FRQYHUWHU
Figure 5.38: A typical 16QAM modulator using two balanced analog
multipliers.
Problem 5.4 The circuit depicted in Figure 5.39 is a balanced AM modulator. Con-
sider the MOS transistors have a threshold voltage VTH and K = 21 µnCox WL , furthermore
vS = VS cos (ωmt) and vRF = VRF cos (ωRFt).
1. Determine the modulation index of the current source, and then find an expres-
sion for the output voltage in terms of the input voltages, vs and vRF . Assume
VRF VGS −VTH and the DC voltage of the gate source junction is VGS0 .
2. If the RF voltage is sufficiently large to completely switch the differential pair,
find an expression for the output in this case.
5.11 Problems 253
VDD
R L C C L R
- Vout +
+
vRF
-
iS
RS
vS
-VDD
Problem 5.5 The circuit depicted in Figure 5.40 shows an envelope detector for
AM signals. The input voltage is vin = 4V (1 + 0.8 cos (2π fmt)) cos (2π f0t) where
fm = 10 kHz and f0 = 1 MHz.
1. Find the spectrum of the output signal near 10 kHz, 1 MHz, 2 MHz, and 4 MHz
through computer simulation.
2. What consideration must be taken into account for the values of R and C to have
minimum distortion?
Ge
V0=0.2 Vout
+
vin R 1.2kΩ C 13.3nF
-
9&&
5 & & 5
4 4 4 4
/2
/LPLWHU
4 56 4
YLQ
+LJK
JDLQ$PS *$ ,66 ,66
9&&
Problem 5.7 Determine the output signal in the circuit of Figure 5.42 where the
inputs are V1 = 20mv f (t), and V2 = 700mv cos 2π × 107t for two cases:
1. f (t) is a normalized analog voice signal of 10 kHz bandwidth limited to ±1.
What kind of modulation is realized in this case?
2. f (t) is a digital signal of 20 kbit/s rate, varying between ±1. What kind of
modulation is realized in this case?
9
9RXW 9RXW
/ ȝ+
& S)
Nȍ
4 4 4 4
ȍ
9
4 4
ȍ
9
$OOVLOLFRQ
WUDQVLVWRUVZLWK P$
Į 9
Problem 5.8 In the amplitude detector depicted in Figure 5.43 for the given input
voltage, determine the output voltage.
Hint: Note that the input amplifier stage (given the unbypassed emitter resistor RE )
operates in the linear region. Furthermore, the envelope detector loads the output-tuned
circuit at the RF frequency by a value of ZL ( jω0 ) = R0 /2.
Ge
M12 V0=0.2 Vout
12V
R1 R0 C0
L1 L2 4kΩ 400pF
46kΩ 2kΩ 2.5nF
60nF
Si
α≈1 ZL(jω0)
Rs
50Ω R2 RE
+ L2=633nH
10kΩ 1.2kΩ
vs M12=1.9μH
L1=5.7μH
vS=1.2v(1+0.5cos105t)cos(2π×107t)
Amplifiers and limiters are among the important building blocks of communication
circuits. In this chapter, we deal with two different types of amplifiers which are
limiting amplifiers and automatic gain controlled amplifiers. The former category has
a high gain and we have introduced a few of its applications in Chapter 5; however,
the latter is concerning methods to control and change the gain of amplifiers which is
quite useful in radio receivers.
and whose gain is reduced with increasing amplitude of the input. The characteristics
of a typical limiter are shown in Figure 6.1.
As Figure 6.1 suggests, for small-signal input, the limiter has merely a linear
response; however, while the input enters the large-signal regime, the output amplitude
will be then limited to a certain value.
As an example, for a differential pair limiter, one could write
It is obvious here that for small values of vi , the output would be linearly related to the
input and for large values of vi , the output will be saturated to a voltage of αRL IEE /2.
9&&
9R
5 5
9R
L& L&
9,3 9,1 9L
,(
9R ,QFUHDVLQJ
LQSXW
DPSOLWXGH
9L
Case II(n = 3): The gain of an amplifier with a single dominant pole, in decibels,
as a function of frequency can be expressed as
A0
AdB (ω) = 20 log r 2 (6.4)
1 + ωωC
In this case, the GBW of each stage is decreased approximately to one-fifth with
respect to the first case, which is of great interest.
Note that here the expression for the frequency response, in case II, would be
of the following form
260 Chapter 6. Limiters and Automatic Gain Control
03
A0
AdB (ω) = 20 log s (6.6)
2 3
1 + ωωC
0
where A0 is the DC gain of each stage (here, 10 dB).
The question which may arise is that is it possible to increase the number of stages
to achieve the same total GBW by lower GBW of each stage? This procedure may
continue till each stage has a gain more than unity. But care must be taken that in the
case of amplifiers, cascading the overall bandwidth would be reduced with respect to
each stage’s bandwidth as demonstrated in the previous example. In a real amplifier,
the nature of frequency response due to finite resistance and capacitance of subsequent
stage will be low-pass. Due to internal feedback at high frequencies and inductive
loads, this behavior of trading gain for bandwidth may change. Thus, increasing the
bandwidth may have an upper bound.
n
4
|Htot ( jω)| =
r
2 (6.7)
1 + ωω0
50 n=4
40 n=3
30
n=2
20
|H|(dB)
10 n=1
0
-10
-20
-30
-40
10 102 103 104
Frequency (MHz)
As Figure 6.3 suggests when the number of stages increases, their overall
bandwidth decreases. Why all the curves pass through the same point at 0 dB
gain?
VOS
VIN VOP
VIP VON
R0
R0 C1 C1
R1 R1
C
VIP VOP
VIN VON
C
Figure 6.5: Offset cancellation loop with negative feedback, where R0 = 300 Ω,
C1 = 0.1 µF, and R1 = 20 kΩ.
262 Chapter 6. Limiters and Automatic Gain Control
Limiting amplifiers are mostly used at the back-end of a receiver and thus are
low-frequency (e.g., 455 kHz) building blocks in communication circuits. The input
impedance in a limiter for a ceramic filter at 455 kHz might be near to 1.5 kΩ or for
a ceramic filter at 10.7 MHz might be 300 Ω. As a result, the matching resistor in
Figure 6.5 is shown to be 300 Ω. Moreover, Figure 6.5 shows the offset cancellation
loop which extracts the undesired DC component at the output and returns a fraction
of it (ideally 1/A of it) with correct sign to the input. Another configuration for offset
cancellation is depicted in Figure 6.6.
Input capacitances in Figure 6.6 are AC short and the 50 Ω impedance at the load
of the error amplifier is for the matching purpose. Thus, the intrinsic output impedance
of the error amplifier must be negligible. However, due to finite output resistance of the
error amplifier, one may decrease 50 Ω resistance to attain a good matching. The loop
mechanism is such that the low-pass filter at the end of the circuit extracts the offset
error voltage and then the error amplifier amplifies the error voltage and returns it to
the input with the opposite polarity. This feedback continues till the DC offset error
reaches zero ideally. Another technique for offset cancellation is shown in Figure 6.7.
As Figure 6.7 shows, the input main amplifier has two differential inputs, the main
input and the auxiliary feedback ones. The error amplifier (A1 ) is placed for the sake
of offset cancellation. Note that the matching criteria is satisfied by 50 Ω brute-force
matching. There are two reasons that the structures shown in Figures 6.6 and 6.7
cannot completely remove the offset voltage. The first reason is the finite gain of the
error amplifier and the second one is their own offset voltages VOS1 for amplifier A1 .
We can easily develop a relation for the overall offset voltage in Figure 6.7 as
q s
0 VOS 2 + A1 2VOS1 2 VOS 2
VOS1 2
V OS = ≈ + (6.8)
AA1 + 1 AA1 A
where in Equation 6.8 offset voltages are considered as random functions with Gaussian
distribution and standard deviation of σ and those are also assumed to be independent.
(UURU
DPSOLILHU
5
$
5 & &
5 5
&
9,1 923
9,3 921
&
Figure 6.6: Offset cancellation loop with active negative feedback, where
R0 = 50 Ω.
6.3 Offset Compensation Circuits 263
(UURU
DPSOLILHU
$
& &
5 5
& 923
9,1 921
9,3
&
5 5 $
Figure 6.7: Offset cancellation loop with negative active feedback and two
differential inputs, where R0 = 50 Ω.
An approximation for Equation 6.8 is valid for the condition of AA1 1. This is also
valid for Figure 6.6, if the gain of the differential input main amplifier is the same for
two paths. Note that the offset voltage of the main amplifier is decreased by the closed-
loop gain; however, unfortunately the offset voltage of error amplifier is just divided
by the main amplifier gain. These gains, however, are low-frequency gains. Thus, if
the offset voltage of the error amplifier is not negligible (i.e., VOS 1 VOS /A1 doesn’t
hold), this component will be dominant. While using high-speed error amplifiers are
not necessary for offset cancellation, we can use larger devices with good matching to
decrease their offset voltage. Finally, note that based on the degree of offset cancellation
needed, we might employ either high-gain error amplifier A1 > 1, or a buffer amplifier
A1 = 1, or a feedback loop without an amplifier.
1
fco2 = (6.12)
2πR0C
Obviously, the overall lower cut-off frequency of this amplifier would be the maximum
of fco1 and fco2 , or
1 + AA1 1
fLF = MAX , (6.13)
2πR1C1 2πR0C
Regarding the amplifier in Figure 6.6 as there is a voltage division at the output of the
error amplifier, A1 is replaced by A21 in Equation 6.10, and therefore, the lower cut-off
frequency in this amplifier becomes
( )
1 + AA2 1 1
fLF = MAX , (6.14)
2πR1C1 2πR0C
The above results suggest that for proper operation of the circuit, we should choose
1/ (2πR1C1 ) very lower than the required fLF . For instance, with a loop gain of 100,
and if we need a lower cut-off frequency of 250 kHz, then the loop bandwidth should
be lower than 2.5 kHz. It is possible to decrease the loop bandwidth by employing a
Miller capacitance in the feedback to obtain a large capacitance, as C1 = (A1 + 1)CF ,
at the input of the error amplifier (Figure 6.8). Like any feedback structure, we
need to examine the stability condition for both the differential and the common-mode
operation. Fortunately, regarding the open-loop gain, the dominant pole of the feedback
circuit is sufficiently near the origin and it is far from the dominant poles of the main
amplifier; as such, it doesn’t pose any challenge to the stability condition.
Cf
50Ω
A1
50Ω
R1 R1
C Cf
VIN VOP
VIP VON
C
Figure 6.8: Offset cancellation loop with negative feedback and use of a Miller
capacitance (instead of a large capacitance) in the feedback amplifier.
6.4 Automatic Gain Control 265
VIN VOP
VIP VON
VAGC
C R
VREF
VDD VDD
R R R R
VON I0/2 I0/2 VOP VON VOP
VAGC
VIP VIN
VIP VIN
(a) ( b)
Figure 6.10: Variable-gain stage with gm variation, (a) Through tail current
variations, and (b) Through second gate voltage variations.
voltage. To solve this problem, we can add two current sources from the outputs to
the ground. By this modification, the current reduced from the load is injected again
to make the common-mode level constant. The constant current is selected to be I0 /2
for both loads. The main drawback of this circuit is the decrease in the output voltage
swing of the circuit for lower gains. Moreover, since the gain is decreased, from input
referred noise point of view, it may result in lower SNR. It can be asserted that with
this technique, the amplifier bandwidth will be constant. Another method to decrease
the gm of the transistor is to push it in a triode region, which in fact lowers its output
impedance and consequently its gain. This can be implemented by a cascode structure
shown in Figure 6.10(b). The gate voltage of the cascode device is controlled by VAGC .
The gain and the cascode devices can be implemented via a single structure MOSFET
as dual-gate. In this method where the current source is not changed, the common-
mode level stays unchanged. However, the input dynamic range will be decreased.
Moreover, this circuit can show an extreme nonlinear behavior for large-signal inputs.
VDD VDD
R R R R
VON VOP VON VOP
VIP VIN
VIP RAGC VIN
(a) (b)
Figure 6.11: Variable-gain stage with (a) load resistor, and (b) series feedback.
one current source implementation has the merit of eliminating the common-mode
noise whose elimination is of interest. The feedback resistor can be implemented by a
MOSFET operation in the triode region. This structure has a loosely fixed bandwidth
and a fixed common-mode voltage. Interestingly, decreasing the gain results in an
increment in the input dynamic range. Moreover, the degeneration resistor (RAGC )
improves the linearity of the amplifier.
s1 s′1
g1
s2 s′2
g2
Input Output
sn s′n
gn
VI VI VIP
VO VO
C I C I
VIN
VI VIN VIP VO
VO
VD VD
t t
(a) (b)
Figure 6.13: (a) Single-ended peak detector and (b) Full-wave peak detector.
6.5 Amplitude Detectors 269
9&& 9&&
9, 9,3
9%%
4
4
9, 4 4 4 4
9,1
92
92
/RZ
& ,( EOHHGLQJ &
FXUUHQW
D E
where VC is the carrier amplitude, VD is the diode cut-in voltage, and m is the mod-
ulation index. Using differential pair transistors, one can detect AM signals with an
amplitude less than VD (why?). The implementation of these peak detectors is shown
in Figure 6.14. In Figure 6.14(a), given the differential pair amplifier, one can detect
small-signal AM inputs, while in Figure 6.14(b), full-wave detection is realized. In
these two implementations, again a constant bleeding current is employed across the
charging capacitor for AM detection. Care must be taken that in these two circuits
a DC bias voltage is required at the input. Furthermore, in Figure 6.14(a), the input
bias should have approximately the same value as the output DC voltage, and the DC
bias of the base of the Q3 should be approximately 2VD +VDC,O . The discharge time
constant of the AM detection in this circuit is approximately
βVC
τ0 = C (6.17)
IE
where β is the current gain of Q4 and VC is the carrier voltage amplitude.
In the AM detector circuit depicted in Figure 6.14(b), a bias voltage larger than VD
is required at the input and the discharging time constant at the output is approximately
τ0 = C VIC , where VC is the carrier voltage amplitude. As mentioned earlier, the
advantage of these circuits with respect to the circuits of the Figure 6.13 is that both
of them can operate with an AC input signal voltage amplitude of a fraction of VD .
Note that the input differential signals should have a low offset voltage to operate
properly.
9,1 923
9,3 921
4 4 4 4 4 4 4 4
5HFHLYHGVLJQDOVWUHQJWKLQGLFDWRU566,
2XWSXWVLJQDO
RSSI
(dB Volts)
Signal stage
gain in dB
RF (dBm)
Figure 6.16 shows the typical output of the indicator versus the input RF power.
As Figure 6.16 suggests, for low-power input signals, merely the final stages sense the
power; however, when the input signal increases, the initial stages sense the power
as well.
VCC
RF RF
QA QB QB′ QA′
VAGC
RE RE
I0
VCC
RF RF
QA Q′A
VAGC RE RE
I0
stage; interestingly due to crossing of the outputs in the Gilbert cell, the common mode
is eliminated. If the control voltage VAGC is high, the currents of transistors Q1 and Q01
flow through the transistors QA and Q0A and a gain of A = RF /RE is obtained. However,
if the control voltage is small, the output current will tend to have small values and the
Gilbert cell output will tend to be zero and a very small gain will be observed. In a
special case where VAGC = 0, the currents totally cancel out each other and a zero gain
is obtained by the assumption of a complete match between the upper tree transistors.
In this circuit, the degeneration resistors RE are used to decrease the low-frequency
gain and increase the dynamic range through the lower tree. The capacitor CE is used
to bypass the degeneration resistors at high frequency and consequently increase the
gain at higher frequencies. The whole scenario is described by Equation 6.18.
272 Chapter 6. Limiters and Automatic Gain Control
vin vAGC
tanh f fcutoff
R
E 2VT
∆IEE = (6.18)
vin vAGC
I0 tanh
tanh f > fcutoff
2VT 2VT
9''
5 5
4 4
$*&
FRQWURO
1 gm
fT = (6.21)
2π Cbe +Cbc
1 gm 3 µn
fT = ≈ (VGS −VTH ) (6.23)
2π Cgs +Cgd 4π L2
where in Equation 6.23, Cgs is the gate–source capacitance, Cgd is the gate–drain
capacitance, µn is the electron mobility, and L is the gate length. For high-speed
operation, NMOS transistors are preferred due to their better mobility. Moreover, the
shorter the length of the device channel and the higher the overdrive voltage the higher
unity current gain frequency, fT , would be obtained. We can also write the maximum
oscillation frequency of a MOS device as
s
1 fT
fmax = (6.24)
2 2πRgCgd
where in Equation 6.24, Rg is the gate resistance. In this equation, we have neglected
the effect of output resistance of the device due to channel length modulation. To
increase the maximum frequency of the transistor, we should increase fT and decrease
Rg and Cgd .
274 Chapter 6. Limiters and Automatic Gain Control
% 4 JP9LQ
9LQ
4 % %഻
9LQ 9LQ
4 9&0
9LQ 9LQ
9LQ
( (
D E
VDD
VO/Vi
L L=RLCL
2
2
gmRL L=0.4RLCL
√2 gmRL L=0
R 2 BW
Vi VO
1.7×BW
f
The inserted inductor is for canceling out the effect of the output capacitance. This
technique is called inductive peaking because it results in a peaking in the frequency
response of the output voltage. For instance, by choosing the value of the inductor as
L = 0.4R2CL , the bandwidth will be increased by 70% with respect to the pure resistive
0
load, i.e., BW = 1.7 BW. The optimum value of the inductor from the bandwidth
point of view doesn’t lead to peaking indeed; however, inductive peaking alleviates the
negative effect of multistage amplifiers in terms of bandwidth. It can be shown that by
increasing the value of the inductor, the bandwidth will be increased and eventually
reaches to its optimum point. If we continue to increase the inductor value, we will
not attain the required bandwidth efficiency anymore and a peaking in the response
will occur which makes the response nonflat. Regarding the inductor model itself and
its self-resonance frequency, as a rule of thumb, the inductor should be chosen such
that its self-resonance frequency is at least twice the cut-off frequency of the amplifier.
Moreover, it is possible to replace the bulky spiral on-chip inductor with its active
counterpart, however, at the cost of higher noise. Figure 6.22 shows the inductor model
with its active counterpart.
Given the equivalent circuit model in Figure 6.22(b), the admittance of the active
inductance circuit can be written as
gm + jCg ω
I =V (6.25)
1 + jRgCg ω
L RS Rg
Rg Rg/ωT
2CP 2CP
= + gmVgs =
Cg Vgs
- 1/gm
(a) (b)
Figure 6.22: (a) Passive (spiral) inductor model. (b) The active MOS inductor
topology and its equivalent circuit.
276 Chapter 6. Limiters and Automatic Gain Control
RgCg ω 1 (6.27)
1 RgCg ω 1 Rg
Z= +j = +j ω (6.29)
gm gm gm ωT
RC RC
Rb Rb
Cin Cin
Cbe Cbe
BJT BJT
RE CE
(a) (b)
bandwidth. The side effect of series feedback is lowering the low-frequency gain
of the amplifier by a 1 + gm RE factor. However, it is possible to increase the load
resistance proportionally to maintain the gain of the amplifier as unchanged. However,
this may result in lowering the output pole frequency due to larger resistance of the
load along with parasitic capacitances. If this pole is not compensated, this would
lead to decrease in the amplifier bandwidth. It is instructive to note that if we choose
CE = 1/ (2π fT · RE ), the emitter capacitance produces a zero in the frequency response
which neutralizes the pole in the response and maintains the amplifier bandwidth. For
example, by insertion of a resistor RE = 100 Ω and a transconductance of 40 mS, the
input capacitance will be lowered by a factor of 5 and we obtain Cin = 34 fF which
results in a cut-off frequency of 39 GHz for the input low-pass response which is
higher than the unity current gain frequency. Now, to maintain the gain, we should
multiply the output resistance by 5, and to neutralize the high-frequency pole of the
emitter, we choose CE = 50 fF. It is possible to increase the emitter capacitance to
achieve more bandwidth which is called emitter peaking. The degeneration resistor
has a lot of advantages, in addition to lowering the input capacitance and increasing
the circuit bandwidth by moving the input pole farther, which are (1) precise gain
control with the ratio of resistors as A = − (RC /RE ) with the criteria of RE 1/gm ,
(2) increasing the input resistance, (3) improving the circuit linearity for large-signal
input, and (4) increasing the dynamic range for a differential stage. In the MOS design,
this issue is not of great concern due to the low intrinsic gate resistance provided by
a proper layout. However, in a similar way to bipolar devices, this resistor can be
employed for the gain control, the amplifier linearity, decreasing the input capacitance,
and increasing the bandwidth with peaking in the source.
300Ω
Figure 6.24: The possible feedback leakage path and the oscillation issue in
multistage amplifiers.
signal to the input will have an amplitude of 20 µv. Thus, for the signals that have an
amplitude greater than 20 µv, the input signal captures the limiting amplifier circuit;
however, if this signal is less than the mentioned value, the returned signal captures
the circuit itself. Most of designers assume this issue is due to poor noise figure;
however, this may happen due to improper layout that worsens the feedback leakage
phenomenon. Furthermore, if the total gain of the amplifiers chain is greater than
80 dB, there would be a possibility of oscillation (if the feedback phase is constructive).
Figure 6.24 illustrates this issue. As it is seen here, there is a possibility of feedback
leakage through either the supply voltage line or the ground line to the input. This
regenerative feedback is one of the pitfalls of designing a limiting stage and thus proper
layout and isolation must be taken into account.
6.9 Conclusion
In this chapter, we discussed different types of limiter circuits. The limiter circuit is a
nonlinear circuit which has a high gain for a limited range of the input signal. When the
signal starts to become large, this circuit limits the signal amplitude to an upper value.
DC offset voltage is another intricate problem in an electronic circuit design. Since the
limiters provide a high gain, offset cancellation techniques are required. We discussed
different types of offset cancellation loops in this chapter. Limiter circuits are used
in FM applications to remove the unwanted amplitude modulation. The AGC in an
amplifier chain is another method to increase the dynamic range of the receiver. The
AGC control loop mainly consists of an amplitude detector plus a negative feedback
bias loop. Furthermore, analog multiplier circuits can be devised in such a way that
it realizes the AGC. The problem of bandwidth enhancement in the RF/IF amplifiers
was presented in this chapter as well. Different techniques were presented in this
6.10 References and Further Reading 279
regard, including fT doubling, inductive peaking, and input capacitance reduction. The
problem of signal leakage and the oscillations due to the feedback path was discussed
as well.
Transmission lines (T-lines) bridge the gap between the field and wave analysis,
on the one hand, and circuit analysis on the other. This makes transmission line
theory an integral part in understanding microwave and mm-Wave devices and circuits.
As it will be seen throughout this chapter, wave propagation through the T-lines
can be formulated by extending the circuit theory basics and making use of some
specific solutions of the Maxwell equations. In this chapter, we provide a profound
understanding of circuit equations governing T-lines using differential equations. The
readers are encouraged to refer to Ref. [1] for a more comprehensive account of the
microwave theory. T-lines play a very important role in modern wireless circuits and
systems which find applications in antenna interfacing to TRX, impedance matching
in mixers and amplifiers, resonator in oscillators and filters, etc.
/1$
'HPRGXODWRU
)UHTXHQF\
V\QWKHVL]HU
namely, capacitors and inductors, tend to have a larger effect, the likelihood of a
positive feedback also increases. For example, consider Figure 7.2, where the model
of a high-frequency amplifier is depicted. As it is seen in Figure 7.2, the matching
circuit elements could be set such that the impedance seen from the antenna and that
of the load are both equal to 50 Ω. It is possible that the circuit could oscillate because
of the existence of parasitic Cgd or Ls , both of which may cause unexpected feedback
at higher frequencies.
/J &JG UG
/6
*DLQRID
7/LQH
I
,GHDO7/LQH
3KDVHRID
1RQLGHDO7/LQH
7/LQH
I
T-line falls with frequency, furthermore its phase changes in a nonlinear manner
with frequency, and therefore it has a variable delay with frequency which causes
phase distortion (or dispersion). Furthermore, a characteristic impedance is defined
for a T-line. As the operating frequency increases, the circuit dimensions become
comparable to the carrier wavelength, the wave behavior of the electromagnetic waves
should be taken into consideration rather than using the lumped element KVL and
KCL relations. As suggested by the maximum power transfer theorem, in the case of
matched terminations, the signal will be completely absorbed by the load. However, as
we will see shortly, if the circuit suffers from a nonzero reflection coefficient, a portion
of the signal, and hence the power, is reflected back.
Capacitor Inductor
50Ω
L L
50Ω 50Ω
L 50Ω L 10Ω
(a) (b)
Open
circuit Z02/ZL
L=λ/4 L=λ/4
50Ω 50Ω
L=λ/4 ZL
L=λ/4
(c) (d)
Figure 7.4: Impedance transformation property of the T-line, (a) The line
terminated by 50 Ω exhibits an impedance of 50 Ω throughout the line, (b) The
line terminated by an unmatched load can exhibit both an inductive or capacitive
impedance seen through it depending on the position, (c) Transformation of a
short circuit to an open circuit using a quarter wavelength T-line, and (d) Load
impedance inversion using a quarter wavelength T-line.
286 Chapter 7. Transmission Lines and Impedance Matching
Figure 7.5: The lumped model of a differential transmission line for a differen-
tial length ∆Z.
∂ 2v
∂ ∂i
= −L (7.7)
∂ z2 ∂ z ∂t
Changing the order of differentiation in Equation 7.7 and using Equation 7.6, we obtain
∂ 2v ∂ 2v
= LC (7.8)
∂ z2 ∂t 2
which corresponds to
∂ 2v
2
∂ v
2
− LC =0 (7.9)
∂z ∂t 2
This is a simple one-dimensional wave equation for the voltage on the line. It is left to
the reader to find a similar equation for the current. Now, for the sake of simplicity,
we assume that the input RF voltage is sinusoidal, having a phasor representation of
v(z,t) = Re{V (z)ejωt }. It follows that
d 2V
− LC(−ω 2 )V = 0 (7.10)
dz2
which can be rewritten as
d 2V
= − ω 2 LC V
(7.11)
dz2
288 Chapter 7. Transmission Lines and Impedance Matching
d 2V
= (ZY )V (7.12)
dz2
Let’s define γ 2 = ZY , where γ is the propagation constant, and then find the solution to
Equation 7.12 as
V (z) = Ae−γz + Be+γz (7.13)
The constants in Equation 7.13 can be found using the initial conditions. The first term
to the left represents the wave prorogating in the positive direction of the z-axis, and
the second term to the left represents the wave prorogating in the negative direction of
the z-axis. Therefore, a more meaningful representation of Equation 7.13 would look
like
V (z) = V0 + e−γz +V0 − Be+γz (7.14)
where, the propagation constant, γ is given by
p √
γ = jβ = ( jωL)( jωC) = jω LC (7.15)
One can observe from Equation 7.15 that γ is frequency dependent. Now, to obtain
the current wave, we substitute V (z) from Equation 7.13 in Equation 7.5, and hence
obtain
jβ + −jβ z 1 + −jβ z
I(z) = V0 e −V0 − e+jβ z = V0 e −V0− e+jβ z (7.16)
jωL Z0
where
r
ωL L
Z0 = = (7.17)
β C
Here, Z0 is the transmission line’s characteristic impedance. This impedance always
shows the quotient of incident voltage to incident current or the quotient of reflected
voltage to reflected current traveling along the transmission line. The voltage traveling
wave equation can be easily obtained by assuming V0 + = |V0 + | ∠φ and arriving at
v+ (z,t) = V0 + cos(ωt − β z + φ + )
(7.18)
The same equation as that derived in Equation 7.18 can also be obtained for the
reflected wave, by means of which the total voltage waveform is given by
v(z,t) = V0 + cos(ωt − β z + φ + ) + V0 − cos(ωt + β z + φ − )
(7.19)
Now, we introduce two new quantities. The distance between two successive planes in
z-direction having the same phase is defined as the wavelength, that is, β (z2 − z1 ) = 2π,
where z2 = z1 + λ and therefore
2π
λ= (7.20)
β
7.2 Wave propagation Equations in Transmission Line for R = 0 and G = 0 289
The wave’s phase velocity is also defined assuming ωt − β z = cte as velocity with
which a specific point on the wave front travels. That is
dz ω 2π f
υp = = = 2π = λ f (7.21)
dt β λ
As the length of the sample line approaches zero, differential forms for Equations 7.25
and 7.26 are obtained as
∂v ∂i
= −Ri − L (7.27)
∂z ∂t
and
∂i ∂v
= − Gv −C (7.28)
∂z ∂t
Now taking a derivative of Equation 7.27, we obtain
∂ 2v
∂i ∂ ∂i
= −R − L (7.29)
∂ z2 ∂z ∂ z ∂t
∂ 2v ∂ 2v
∂v ∂v
= −R −Gv −C − L −G −C (7.30)
∂ z2 ∂t ∂t ∂t 2
∂ 2v
2
∂v ∂ v
− (RG) v − (RC + LG) − LC =0 (7.31)
∂ z2 ∂t ∂t 2
d 2V
− (RG)V − (RC + LG) jωV − LC(−ω 2 )V = 0 (7.32)
dz2
Equation 7.32 can be rewritten as
d 2V
= (RG)V + jω(RC + LG)V − ω 2 LC V
2
(7.33)
dz
Defining Z = R + jωL and Y = G + jωC, Equation 7.32 can be written as
d 2V
= (ZY )V (7.34)
dz2
The constants in Equation 7.35 can be computed using the initial conditions. The first
term to the left represents the wave prorogating in the positive direction of the z-axis,
7.3 Characteristic Impedance of a Line 291
and the second term represents the wave prorogating in the negative direction of the
z-axis. Therefore, a more meaningful representation of Equation 7.35 would look like
V (z) = V0 + e−γz +V0 − e+γz (7.36)
It also follows that
p
γ = α + jβ = (R + jωL)(G + jωC) (7.37)
where the real and imaginary parts describe the attenuation and the phase constants,
respectively. Here the propagation constant has a nonlinear frequency dependency, and
therefore the transmission line is dispersive. The current waveform can be obtained by
substituting Equation 7.36 in the phasor form of Equation 7.27 and we obtain
γ
V0 + e−γz −V0 − e+γz
I(z) = (7.38)
R + jωL
As before, the wave traveling in the positive z-direction is
v+ (z,t) = V0 + e−αz cos(ωt − β z + φ + )
(7.39)
And hence the overall voltage waveform at the input is given by
v(z,t) = V0 + e−αz cos(ωt − β z + φ + ) + V0 − e+αz cos(ωt + β z + φ − ) (7.40)
The phase velocity can be defined as before for the lossy case, here we have
ω
υp = np o (7.41)
Im (R + jωL)(G + jωC)
nH fF
Example 7.1 Consider Equation 7.45 for a line with L = 0.273 mm , C = 93.5 mm ,
mΩ µS
R = 170 mm , and G = 60 mm . Derive an expression for
the phase and amplitude
of
the line characteristic impedance from 0 to 1 GHz. λ` = 2n−2 4 (n ≥ 1) , and plot
the real and the imaginary parts as a function of frequency.
Solution:
It follows from Equation 7.45 that
s s
R + jωL 0.17 + jω(0.273)10−9
Z0 = = (7.46)
G + jωC 60 × 10−6 + jω(93.5)10−15
,PDJLQDU\
SDUWȍ
SDUWȍ
5HDO
)UHTXHQF\*+]
As it can be seen in the above figure, Z0 becomes almost pure real and inde-
pendent of frequency at the higher portion of the spectrum.
The current waveform as a function of the line characteristic impedance can also be
written as
1
V0 + e−γz −V0 − e+γz
I(z) = (7.47)
Z0
will no longer experience any loss and will experience only a phase shift. The line
characteristic impedance becomes in this case as
s r
0 + jωL L
Z= = (7.49)
0 + jωC C
As it is observed from Equation 7.49, the line characteristic impedance becomes purely
real and it is also independent of frequency. The voltage and the current waveforms
can also be derived in the lossless case as
V (z) = V0 + e−jβ z +V0 − e+jβ z (7.50)
and
V0 + −jβ z V0 − +jβ z
I(z) = e − e (7.51)
Z0 Z0
As an example, for the values given in Example 7.1, the phase velocity amounts to
1.98 × 108 m/sec. If the medium in between the two conductors is homogeneous with
the permittivity ε and the permeability µ, it can be shown that for the transmission
line, one can write
LC = µε (7.53)
2π 2π
λ= = √ (7.54)
β ω LC
We also know that the speed of TEM electromagnetic propagation wave in a dielec-
tric/magnetic medium is
1 c
υg = √ =√ (7.55)
µε µr εr
where c is the speed of light in vacuum. Using Equation 7.53, we end up with υP = υg
which means the phase velocity in a TEM transmission line is the same as the speed of
propagation in a free space medium.
I(z)
+
V(z) Z0,Ɣ ZL
-
Z
l Z=0
end. The voltage at the source end of the line is of interest, which can be written from
Equation 7.36 as
where ` is the distance from the load and the first term in the right-hand side is the
traveling wave in the positive z-direction and the second term is the reflected wave
in the negative z-direction. The current waveform can also be derived in a similar
manner as
V−
V (−`) = V0+ eγ` +V0− e−γ` = V0+ eγ` 1 + 0+ e−2γ` (7.58)
V0
where V0 + is the traveling wave vector toward the load at Z = 0 and V0 − is the reflected
wave from the load at Z = 0. The ratio of the reflected wave to the incident wave at the
load is defined as reflection coefficient and can be computed as
V0− e−γ`
ΓL (−`) = = ΓL (0)e−2γ` (7.59)
V0+ e+γ`
Using the above definition, the voltage and the current waveforms can be written using
the reflection coefficient as
and
V0+ γ`
I (−`) = e (1 − ΓL e−2γ` ) (7.61)
Z0
7.4 Terminated Transmission Lines 295
It is easy to compute the input impedance by finding the ratio of the voltage to the
current phasors as
1 + ΓL e−2γ`
V (−`)
Z (−`) = = Z0 (7.62)
I (−`) 1 − ΓL e−2γ`
which gives the input impedance of an arbitrarily terminated T-line. If one chooses
` = 0, the load impedance is given as
1 + ΓL
Z (0) = Z0 = ZL (7.63)
1 − ΓL
ZL − Z0
ΓL = (7.64)
ZL + Z0
Factoring Z0 and ZL out, Equation 7.67 can be rewritten using hyperbolic functions as
ZL cosh (γ`) + Z0 sinh (γ`)
Z (−`) = Z0 (7.68)
Z0 cosh (γ`) + ZL sinh (γ`)
Now, we derive the above expressions for a lossless T-line. The voltage and the current
waveforms are given by
V (−`) = V0+ ejβ ` 1 + ΓL e−2jβ ` (7.70)
and
V0+ jβ `
I (−`) = e 1 − ΓL e−2jβ ` (7.71)
Z0
Therefore, it follows that
!
1 + ΓL e−2jβ `
V (−`) ZL + jZ0 tan (β `)
Z (−`) = = Z0 = Z0 (7.72)
I (−`) 1 − ΓL e−2jβ ` Z0 + jZL tan (β `)
Since in the lossless T-line tanh (γ`)=tanh ( jβ `)= j tan (β `), by replacing tanh (γ`) in
Equation 7.69, we can obtain the same result as above. It follows from Equation 7.72,
the input impedance has a periodic property as a function of distance, and we have
2π nλ
β ` = nπ ⇒ ` = nπ ⇒ ` = (7.73)
λ 2
This means that the input impedance attains the same value for each λ2 length.
Equation 7.72 is used for a lossless transmission line while Equation 7.69 is used for a
lossy transmission line. The calculation of these complex equations can be simplified
through the use of the Smith chart which is provided in section 7.10.
Example 7.2 For the circuit in Figure 7.9, compute the input impedance and the
reflection coefficient at the input and at the load side.
(a) First, consider the following parameters for the line and find the results at 1 GHz.
(b) Consider the line is lossless and repeat part (a).
pF S
L = 1.125 µH Ω
m ,C = 450 m , R = 5 m , G = 0.01 m
Zin
l =0.5λ
Solution:
(a) We have
p
γ = α + jβ = (R + jωL)(G + jωC) (7.74)
q
5+ j(2π ×109 )×1.125×10−6 0.01+ j(2π ×109 )×450×10−12
=
1
= 0.3 + j141.371
m
7.4 Terminated Transmission Lines 297
s
R + jωL
Z0 = ≈ 50 Ω (7.75)
G + jωC
λ 2π
Since ` = 2 and β = λ , then from Equation 7.76, the line length becomes
2π
λ β π
`= = = = 2.22 cm (7.77)
2 2 β
Therefore
30 − j20 + 50 tanh (α + jβ ) βπ
Z (−`) = 50 (7.78)
50 + (30 − j20) tanh (α + jβ ) βπ
As it is seen here the input reflection coefficient is slightly different from the load
reflection coefficient (because the line length is λ2 and the line is lossy).
(b) For the input impedance, we have from Equation 7.72
!
30 − j20 + j50 tan 2π
ZL + jZ0 tan (β `) λ 0.5λ
Z (−`) = Z0 = 50
50 + j(30 − j20) tan 2π
Z0 + jZL tan (β `) λ 0.5λ
= 30 − j20 (7.82)
As it is seen here the reflection coefficient has exactly the same value as the input
(because the line length is λ2 and the line is considered as lossless).
298 Chapter 7. Transmission Lines and Impedance Matching
for the reflection coefficient at the load and the source, we have
ZL − Z0 Z0 − Z0
ΓL = = =0 (7.85)
ZL + Z0 Z0 + Z0
Therefore, for this case, the input impedance is equal to the characteristic impedance
and the reflection coefficient will be equal to zero throughout the line, irrespective of
the location. No reflection will therefore occur on the line and perfect matching has
been achieved, and the voltage and the current waveforms can be derived as
and
V0+ +jγ`
I (−`) = e (7.87)
Z0
It implies that a voltage or a current wave can be moved from the point −` to point zero
with only a phase shift and then they will be absorbed by the load (with no reflection).
It should be noted that the normal PCB circuit wires cannot transmit the high-frequency
signals. Using transmission lines in PCBs, we can transfer very high frequency signals
to any devised distance.
I(-l )
+
Z0,Ɣ V(-l ) ZL
-
As it is obvious from Equation 7.89, the input impedance in this case is always an
imaginary quantity. This suggests that any arbitrary value of reactance can be obtained
with a T-line terminated to a short circuit. The value of the input impedance in this
case is depicted in Figure 7.12.
As it is obvious from Figure 7.12, the impedance assumes the values of zero and
infinity at even multiples of λ4 and at odd multiples of λ4 , respectively. This result shows
I(-l )
+
Z0,Ɣ V(-l )
-
"
_=_ȍ
,QGXFWLYH
EHKDYLRU
" OJ
&DSDFLWLYH
EHKDYLRU
that moving away from a short-ended transmission line first for less than quarter wave
length distances, the input becomes inductive and for a length in excess of a quarter-
wave length, it becomes capacitive. In practice, we use short-ended transmission lines
shorter than a quarter of a wavelength to realize inductive impedances. The reflection
coefficient at the load can be computed from Equation 7.64 as
ZL − Z0 0 − Z0
ΓL = = = −1 (7.90)
ZL + Z0 Z0 + 0
As it is obvious from Equation 7.92, the input impedance in this case is always
imaginary, meaning that any value of a reactive impedance can be realized by this line.
The input impedance in this case is depicted in Figure 7.14.
As it is obvious from Figure 7.14, for odd multiples of λ4 and even multiples of λ4 ,
the input impedance is that of short circuit and open circuit, respectively. This result
shows that moving away from an open-ended transmission line first (for the lengths less
than a quarter wave length), the input becomes capacitive and for a length in excess of a
quarter wave length, it becomes inductive. In practice, we use open-ended transmission
lines shorter than a quarter of a wavelength to realize capacitive impedances. The
reflection coefficient can be computed from Equation 7.64 as
ZL − Z0
ΓL = =1 (7.93)
Z0 + ZL ZL →∞
I(-l )
+
Z0,Ɣ V(-l )
-
"
_=_ȍ
,QGXFWLYH
EHKDYLRU
" OJ
&DSDFLWLYH
EHKDYLRU
I(-l )
ZTH
+ +
VTH Z0=β V(-l ) ZL
- -
Zin l
ZTH
+ +
VTH V(-d) Zin
- -
Figure 7.16: Equivalent model of Figure 7.15 for input impedance calculation.
Equation 7.96 can be written for the input, and using Equation 7.94, we arrive at
Zin
V (−d) = V0+ ejβ d 1 + ΓL e−j2β d = VTH (7.97)
Zin + ZTH
The incident wave phasor can be obtained from Equation 7.96; it follows that
+ Zin −jβ d 1
V0 = VTH e (7.98)
Zin + ZTH 1 + ΓL e−j2β d
Now, by substituting Equation 7.98 in Equation 7.96, the voltage can be found at any
point along the T-line as
!
−j2β `
Zin −jβ (d−`) 1 + ΓL e
V (−`) = VTH e (7.99)
Zin + ZTH 1 + ΓL e−j2β d
It follows from Equation 7.99 that the impedance at distance d from the load as
!
V (−d) 1 + ΓL e−j2β d
Zin = Zin (−d) = = Z0 (7.100)
I (−d) 1 − ΓL e−j2β d
Then, an explicit expression of the ratio of the source impedance divided by the input
impedance can be found as
−j2β d
Zin (−d) Z0 1+Γ Le
1−ΓL e −j2β d
= (7.101)
Zin (−d) + ZTH 1+ΓL e−j2β d
Z0 1−Γ e−j2β d + ZTH
L
which reduces to
1 + ΓL e−j2β d
Zin (−d) Z0
= (7.102)
Zin (−d) + ZTH ZTH + Z0 1 − ΓL e−j2β d ZZTH −Z0
TH +Z0
7.6 Source and Load Mismatch in Lossless Lines 303
Finally, the voltage and current waveforms can be found at any point along the line
using Equation 7.105.
Example 7.3 Consider the circuit shown in Figure 7.17(a). Find the voltage
waveform at the input. (Hint: Derive the input such that it exhibits wave transmis-
sion back and forth toward the load and from the source. This behavior is depicted
in Figure 7.17(b).)
d
ZTH
+
VTH Z0,β ZL
-
(a)
d
ZTH
V
VΓLe-j2βd
VΓLe-j2βdΓS
+ VΓLe-j2βdΓSΓLe-j2βd
VTH ZL
VΓLe-j2βdΓSΓLe-j2βdΓS
-
(b)
Figure 7.17: (a) Demonstration of a T-line with its load and source
impedances, and (b) transmission of the wave back and forth on the line.
304 Chapter 7. Transmission Lines and Impedance Matching
Solution:
First, the voltage is divided between the source impedance and the input impedance.
Then, it moves along the line and a fraction of it is reflected, proportional to the
reflection coefficient, ΓL . This repeats while the wave reaches the sources and
is reflected back proportional to the source reflection coefficient, ΓS . In order to
quantify this behavior, we have
Z0
V (−d) = VTH 1 + ΓL e−j2β d + ΓL e−j2β d ΓS + · · ·
Z0 + ZTH
(7.106)
!
∞ i
Z0
= VTH 1 + (ΓS + 1) ∑ ΓL e−j2β d ΓS i−1
Z0 + ZTH i=1
which results in
Z0 ΓS + 1 1
V (−d) = VTH 1+ − 1 (7.109)
Z0 + ZTH ΓS 1 − ΓL ΓS e−j2β d
Now, we can use Equation 7.109 to derive an expression similar to that derived in
Equation 7.105 for d = `
!
1 + ΓL e−j2β d
Z0
V (−d) = VTH (7.110)
Z0 + ZTH 1 − ΓL Γs e−j2β d
As such, the general Equation 7.105 reduces to Equation 7.110 for ` = d. This
should make it clear that in fact the wave travels back and forth in multiple reflec-
tions to reach a steady-state condition.
The important property revealed by Equation 7.111 is that it can transform the load
impedance to a value proportional to the inverse of load impedance (multiplied by
the square of characteristic impedance of the line). This property finds an important
application in impedance matching networks where one intends to change a real
impedance value to another real impedance value.
As it can be seen, the voltage along the line varies between a maximum and a minimum
value depending on the phase of the second term in the parentheses, i.e., φL − 2β `.
Along the transmission line at the point (points) where this phase is equal to zero, there
will be a maximum voltage and at the point (points) where this phase is equal to π,
there is a minimum voltage. The ratio of these two quantities is called voltage standing
wave ratio (VSWR).
C1 C1
L C2
Inductor Capacitor
ȍ ȍ
6KRUW 2SHQ
FLUFXLW FLUFXLW
D
ȍ ȍ
,PSOHPHQWDWLRQRIDVHULHVLQGXFWRU
E
ȍ
,PSOHPHQWDWLRQRIDSDUDOOHOFDSDFLWRU
F
+
Vmax V (1 + |ΓL |) 1 + |ΓL |
V SW R = = 0+ = (7.113)
Vmin V0 (1 − |ΓL |) 1 − |ΓL |
finds another application in amplifiers, where in order to achieve the maximum power
transmission in both input and the output, matching must be present between the source
and the input on the one hand, and the load and the output on the other. There are quite
a few points which merit specific attention in matching networks:
(1) Complexity: designing the most simple matching network possible is crucial. After
all, a smaller matching network is less expensive, more reliable and suffers from a
smaller attenuation with respect to its more complex counterparts;
(2) Bandwidth: every single matching network can only provide matching in a specific
bandwidth given by its specifications. This is, however, inadequate for many applica-
tions where a wideband matching is desired. There exists a myriad of techniques for
bandwidth extension, which as expected, come at the price of complexity;
(3) Implementation: the designer may prefer one matching structure to another based
on the type of the T-line, lumped elements, or the waveguide. For example, tuning
screws in waveguide arms provide better flexibility than the quarter wave length T-line,
and
(4) Tunability: In some applications, it is desirable to be able to fine-tune the matching
network so as to achieve the maximum power transmission to the load. Some architec-
tures are better fit for this property.
Consider the circuit depicted in Figure 7.20. We will call circuits of this kind which
have a frequency-selective behavior a resonant or a tank circuit thereafter. Before
dealing with the circuit details, let’s first discuss a parameter which provides a measure
of loss in energy storage element. Referred to as quality factor, Q, it is defined as
f0
Q= (7.114)
∆f
wherein f0 is the resonant frequency of the circuit and ∆ f is the corresponding 3 dB
bandwidth. Nevertheless, the presence of load and source impedances degrades the
overall quality factor of the circuit. We call the matching in this case a loaded one.
The effect of the source impedance on the frequency response of the resonant network
is depicted in Figure 7.21. As it is evident in Figure 7.21, when the input impedance
is equal to 50 Ω, the frequency response is wider, while for an input impedance of
1000 Ω, a much narrower response is observed. The inductance of the inductor is also
important in the overall quality factor of the circuit. As depicted in Figure 7.22, with a
higher inductance, a lower Q is obtained mainly due to ohmic losses. In this circuit,
RS
Q = Lω 0
= RSCω0 .
As stated in the beginning of this chapter, one of the properties of matching
network is maximum power transmission. As you can recall from the basic circuit
theory, in a DC circuit, maximum power transmission between the load and the source
RS
+ L
Vin RL C
Rloss
0 RS
Vout
|Vout/Vin|(dB)
-10
-20
-30 Vin XL XC
-40
-50 Q=1.1, RS=50Ω
Q=22.4, RS=1000Ω
-60 1 L=50nH,C=25pF
10 102 103
Frequency (MHz)
Q+ S) ȍ
_=MȦ_
=
Q+ S) ȍ
)UHTXHQF\0+]
=
0.26
0.24 1Ω Vout
0.22
P(Watt)
0.2
0.18
0.16
0.14 Vin RL
0.12
0.1
0.08
10-1 100 101
RL(Ω)
impedances occurs when the two have equal values. Depicted in Figure 7.23 is the
power that is delivered to a variable load for a 1 V source.
Now, let’s consider the more general case depicted in Figure 7.24, which includes
a source impedance having both resistive and inductive components, and a load
consisting of a capacitive and a resistive component. If the capacitor and the inductor
resonate at some frequency, the circuit then reduces to that shown in Figure 7.22,
and maximum power transmission is achieved by choosing equal load and source
impedances. That is the total load impedance should be the complex conjugate of the
7.9 Impedance Matching: The L-Section Approach 309
ZS
RS jXS ZL RS
-jXL
+ RL + RL
Vin Vin
Figure 7.24: The general case of maximum power transmission to the load.
source impedance (ZL = ZS∗ ). In order to achieve this condition, we now proceed to
introduce some popular matching structures, known as L-sections. Different variants
of an L-section are depicted in Figure 7.25.
We analyze, in the final section of this chapter, all the structures depicted in
Figure 7.25 using a powerful tool called Smith chart. Now, we consider the basic
impedance matching methods. See Figure 7.27.
RS L RS L
+ +
Vin C RL Vin C RL
Lowpass Lowpass
RS C RS C
+ +
Vin L RL Vin L RL
Highpass Highpass
ȍ Mȍ
Mȍ ȍ
= Mȍ
-j300Ω
100Ω
Z1
Figure 7.27: The equivalent circuit of the network shown in Figure 7.26 at the
operating frequency.
Firstly assume that the source is terminated to an RL = 100 Ω load and we calculate
the power delivered to this load. Secondly assume that in Figure 7.26, the LC matching
network (L-section) is omitted and the source is terminated to an RL = 1000 Ω load
and we recalculate the power delivered to this load. The ratio of this latter power to the
former power is defined as the mismatch loss.
vmismatch 2
vmismatch 2 Rmatch
pmismatch Rmismatch
L = −10 log = −10 log = −10 log =
pmatch vmatch 2 vmatch Rmismatch
Rmatch
(7.115)
2
0.909vin 100
− 10 log = 4.8 dB
0.5vin 1000
As demonstrated by Equation 7.115, the loss is equal to 4.8 dB. Now, if we employ
the circuit shown in Figure 7.26, the impedance seen at the load, Z1 ,will amount to
100 − j300 Ω. Let’s calculate the equivalent impedance of the RC section as
jXC RL − j333(1000)
Z= = = 100 − j300 (7.116)
jXC + RL − j333 + 1000
This is equivalent to the circuit shown in Figure 7.27 at the operating frequency. Now,
if an inductor of 300 jΩ is added in series to this network, then the load impedance will
be purely resistive and equal to 100 Ω. This is conceptually depicted in Figure 7.28.
What we have learned from the concept of impedance matching thus far suggests that a
parallel element can lower the real part of the impedance level of the load, which for the
+j300Ω
-j300Ω
100Ω = 100Ω
Z2 Z2
case studied here, a 1000 Ω load was lowered to 100 Ω. The series element can resonate
with the other energy-storing element, resulting in a purely resistive impedance equal
to the source impedance. This, however, points to a possible limitation of series or
shunt resonant matching network, which is their narrowband response due to single
frequency matching or resonance.
Rp Xp 2
Rs = (7.119)
Rp 2 + Xp 2
and
Rp 2 Xp
Xs = (7.120)
Rp 2 + Xp 2
which represent the equivalent series and parallel impedances, respectively. Given the
fact that QP = QS , it follows from Equation 7.119 that
Rp
= 1 + Qp 2 = 1 + Qs 2 (7.121)
Rs
XS
Xp Rp
RS
Figure 7.29: The equivalent circuit of a reactance with a limited quality factor.
312 Chapter 7. Transmission Lines and Impedance Matching
Now, for an L-section matching, using two different reactances, one can use the source
resistance, the load resistance, and Equation 7.122, as the starting point and then
one chooses two reactances of the opposite signs to realize the conjugate matching
condition. It is obvious that this matching procedure is applicable once Rp > Rs . This
principal matching procedure is demonstrated in Figure 7.30.
Qs=Xs/Rs
Qp=Rp/Xp
Rs
Xs
+ Xp Rp
Vin
Rp
Qs=Qp= -1
Rs
Example 7.4 Consider the circuit shown in Figure 7.31. Design a matching
network at 100 MHz which matches a source impedance of 100 Ω to the load
impedance of 1000 Ω. Also, you may assume that the circuit is DC coupled.
RS L
+
Vin C RL
Solution:
We first compute the quality factor from Equation 7.122 as
r r
Rp 1000
Qs = Qp = −1 = −1 = 3 (7.123)
Rs 100
7.9 Impedance Matching: The L-Section Approach 313
and
−Rp −1000
Xp = = = −333.3 Ω (7.125)
Qp 3
respectively. The values of the inductor and the capacitor can both be easily
calculated as
Xs 300
L= = = 477 nH (7.126)
ω 2π(100 × 106 )
and
1 1
C=− = = 4.8 pF (7.127)
ωXp 2π(100 × 106 )(333.3)
respectively.
Now that we have learned the impedance matching concept, we turn to complex load
and source impedances. This can be the case while circuits are being interfaced to real-
world impedances like antennas, mixers, T-lines, transistors, and other components,
where their input impedance is both complex and frequency-dependent. A possible so-
lution can be absorption of impedances within the matching network. This can be done
by absorbing the stray capacitances into parallel matching capacitors, and by absorbing
the stray inductances into series matching inductors. We elaborate on this point in
Example 7.5 in which series resonance occurs at the desired frequency.
Example 7.5 Consider the circuit depicted in Figure 7.32. Using impedance
absorption technique, match the source and load impedances at 100 MHz.
100Ω 200nH
Z
match 2pF 1000Ω
100-j126
Solution:
The first step is to consider only the real part of the source impedance which is
100 Ω at 100 MHz and the real part of the load impedance which is 1000 Ω at
100 MHz. Using the numerical results of Example 7.4, the matching network will
look like that depicted in Figure 7.33. The matching of the 1000 Ω to the 100 Ω
source would need a 477 nH series inductor and a parallel 4.8 pF shunt capacitor as
demonstrated in Example 7.4. By subtracting the existing 200 nH inductance from
the 477 nH needed one and subtracting the 2 pF capacitance from the 4.8 pF needed
one, we obtain the resultant values shown in the dashed recangle in Figure 7.33.
Figure 7.33: The proper matching network for complex source and complex
load matching.
Note that here we had positive values for the series inductance and the parallel
capacitance, given the fact that the total needed inductance and the total needed
capacitance in the matching network were larger than the stray inductance and the
stray capacitance, respectively.
Example 7.6 Design a matching network at 75 MHz for the circuit shown in
Figure 7.34. Employ DC blocking.
50Ω
Z
match 40pF 600Ω
Solution:
The desired matching network which employs DC blocking is illustrated in
Figure 7.35.
7.9 Impedance Matching: The L-Section Approach 315
50Ω C
L 40pF 600Ω
1 1
L= = = 112.6 nH (7.128)
ω 2C 6 2
2π(75 × 10 ) × 40 × 10−12
50Ω
Z
match 112.6nH 40pF 600Ω
Now that the stray capacitance has been tuned out, we should match the 50 Ω
source to 600 Ω load resistor. Hence,
r r
Rp 60
Qs = Qp = −1 = − 1 = 3.32 (7.129)
Rs 5
and
Rp 600
Xp = = = 181 Ω (7.131)
Qp 3.32
which corresponds to
Xp 181
L= = = 384 nH (7.132)
ω 2π(75 × 106 )
316 Chapter 7. Transmission Lines and Impedance Matching
and
1 1
C=− = = 12.7 pF (7.133)
ωXs 2π(75 × 106 )(166)
50ȍ 12.78pF
Now, if we replace two parallel inductors with one, the final matching network
will be as depicted in Figure 7.38.
50Ω 12.78pF
Another method which is used for impedance matching is known as the π and T
method, which is illustrated in Figure 7.39.
A π network can be formed by cascading two L-sections with a virtual resistor in
between, as shown in Figure 7.40.
The negative signs in the parallel reactances are just for representation purposes;
however, the important point is that they differ with their series counterparts in sign.
Therefore, if Xp1 is a capacitor, Xs,1 must be an inductor and vice versa. Similarly, if Xp2
RS RS
X1 X3 X2
X2 RL X1 X3 RL
Req
RS
Xs1 Xs2
-Xp1 -Xp2 RL
is a capacitor, then Xs2 must be an inductor and vice versa. Now, from Equation 7.122,
it follows that
s
RH
Q= −1 (7.134)
Req
where RH is the larger resistor on the source and the load side, and Req is the virtual
resistor between the two networks. Evidently, RH should be larger than Req .
Example 7.7 Consider the circuit shown in Figure 7.40. Design a π match-
ing network such that the input impedance of 100 Ω is matched to a 1000 Ω load
impedance. The loaded Q of each L-section is equal to 15.
Solution:
We first consider Q from Equation 7.134, and calculate the virtual resistor as
RH 1000
Req = = = 4.42 Ω (7.135)
Q2L + 1 226
We can then use Equation 7.117, the series and parallel reactances, as depicted in
Figure 7.41, can be calculated as
Req
Xs2
-Xp2 RL
RL 1000
Xp2 = = = 66.7 Ω (7.137)
Qp2 15
Now, we analyze the left-hand L-section. The quality factor for this circuit, as
depicted in Figure 7.42, can be found as
Req
Xs1
RS -Xp1
s r
RS 100
Qs1 = Qp1 = −1 = − 1 = 4.6 (7.138)
Req 4.42
and using Equation 7.117, the parallel and series reactances can be calculated as
and
RS 100
Xp1 = = = 21.7 Ω (7.140)
Qp1 4.6
Now that we have the values of all the reactances, the overall matching network
can be chosen as in Figure 7.43 in a different version depending on the chosen sign
of the reactances.
Req
100Ω
Xs1 Xs2
The main point to remember here is that the reactances in each branch of
L-sections have a different sign. Therefore, the other structures can also be used for
matching, all of which are depicted in Figure 7.45. In total, four combinations of π
section matching network (depending on low-pass/DC coupled or high-pass/AC
coupled L-section being chosen) are possible to realize.
It could be verified through circuit simulations that the overall quality factor
and therefore the corresponding bandwidth of this matching circuit could be found
through the following relation
Figure 7.45: Variants of the matching network with inductors and capacitors
in a π structure.
320 Chapter 7. Transmission Lines and Impedance Matching
RS
Xs1 Xs2
-Xp1 -Xp2 RL
Req
The reader may wonder that which one of the parameters should be chosen for the
matching network. In order to pick one of them, it should be noted that the following
parameters might be taken into account: DC coupling or AC coupling, the values of
stray capacitances or stray inductances to be tuned out (to be deduced from the match-
ing network), harmonic voltages or harmonic currents to be eliminated, the frequency
response of the matching network, and finally realizable values for the matching
elements.
We now proceed by discussing the T matching network. Matching with a T
network is very much similar to its π counterpart and it is done by two L-sections as
well. With the exception that both of the L-sections are matched to a larger value of
virtual resistance and the horizontal part of the L-section, that is, the series reactance
comes in series with the load or the source. As a result, the parallel arms of the two
L-sections will eventually appear in parallel. The T matching network is depicted in
Figure 7.46.
Let’s now return to our definition of Q as
r
Req
Q= −1 (7.142)
Rl
where Req is the virtual resistance to be matched to the load and the source, and Rl
is the minimum of the load and the source resistances. Example 7.8 clarifies this
point.
Example 7.8 Consider the circuit illustrated in Figure 7.46. Design four dif-
ferent matching networks which match a 10 Ω source impedance to a 50 Ω load
impedance. The loaded Q is chosen to be 10.
Solution:
The virtual resistance can be found from Equation 7.142 as
and the values of the series and the parallel reactances, as depicted in Figure 7.47,
can be calculated using Equation 7.117 as
7.9 Impedance Matching: The L-Section Approach 321
RS
Xs1
-Xp1 Req
Now, it follows from Equation 7.142 that the quality factor of the right-hand
L-section, as depicted in Figure 7.48, is
RL
Xs2
Req -Xp2
r r
Req 1010
Qs2 = Qp2 = −1 = − 1 = 4.4 (7.146)
RL 50
and
Req 1010
Xp2 = = = 230 Ω (7.148)
Qp2 4.4
Finally, the four possible T matching networks are depicted in Figure 7.49. Note
that the final parallel reactance is the resultant parallel combination of the two
middle reactances of each L-section.
322 Chapter 7. Transmission Lines and Impedance Matching
It could be verified through circuit simulations that the overall quality factor
and therefore the corresponding bandwidth of this matching circuit could be found
through the following relation
10Ω
Xs1 Xs2
As in the previous example, here each one of the four variants could be chosen ac-
cording to the following criteria: DC coupling or AC coupling, the required frequency
response, low-pass or high-pass response, the requirement to deduce the stray capac-
itances or inductances from the matching network, harmonic voltages or harmonic
currents to be eliminated, and finally realizable values for the matching elements. Thus
far, we have learned to match any passive load to any passive source impedance using
L, π, or T networks. We should have realized in this section, that matching networks
are indispensable parts of any radio-frequency circuit. In the next section, we introduce
a powerful tool which could be used for matching networks calculations, that is the
Smith chart application.
Now, if we define ZL /Z0 as the normalized impedance and writing ΓL in terms of its
real and imaginary parts, it follows that
Zn − 1 R + jX − 1
Γr + jΓi = = (7.151)
Zn + 1 R + jX + 1
In order to find an explicit expression for the Smith chart contours, we rewrite
Equation 7.151 as
1 + ΓL
Zn = (7.152)
1 − ΓL
We proceed by replacing the real and the imaginary parts of the reflection coefficient
and the normalized impedance in Equation 7.152, hence,
1 + Γr + jΓi
R + jX = (7.153)
1 − Γr − jΓi
By multiplying the right-hand side of Equation 7.153 by the complex conjugate of the
denominator, we arrive at
Equating the real and the imaginary parts of Equation 7.154, we have
1 − Γr 2 − Γi 2
R= (7.155)
(1 − Γr )2 + Γi 2
and
2Γi
X= (7.156)
(1 − Γr )2 + Γi 2
Equations 7.155 and 7.156 play a crucial role in finding an expression for the Smith
chart contours. Rewriting Equations 7.155 and 7.156 in two complete square forms in
terms of Γr and Γi , we arrive at
2 2
R 1
Γr − + Γi 2 = (7.157)
1+R 1+R
and
1 2
2
1
(Γr − 1)2 + Γi − = (7.158)
X X
Equations 7.157 and 7.158 represent constant resistance and constant reactance con-
tours in the Γr -Γi plane. Each point in this plane corresponds to a unique reflection
324 Chapter 7. Transmission Lines and Impedance Matching
Γi Γi
Inductive reactance component
+j 1
0 0.2 0.5 1 2 5 Γr 0
Γr
-j 1
Capacitive reactance component
Constant Constant
resistance circles reactance circles
Figure 7.50: Constant-resistance and constant-reactance circles within |Γ| ≤ 1
plane.
coefficient and therefore a unique impedance in that plane. Equation 7.157 represents a
group of circles centered at (x, y) = (R/ (1 + R) , 0) and with a radius of r = 1/ (1 + R).
Equation 7.158 represents a group of circles centered at (x, y) = (1, 1/X) and with
a radius of r = 1/X. Drawing the two groups of circles for loads having a positive
real part (within a circle centered at the origin and with the unity radius) results in
the Smith chart. The intersection of the reactance and the resistance circles results in
a specific impedance in the Γr -Γi plane. Constant-resistance and constant-reactance
circles are shown in Figure 7.50.
Now, we mention a few points regarding the Smith chart. (1) The intersections of
constant-resistance circles with the Γr axis (x-axis) are purely real impedance (R + j0)
points. (2) The intersections of constant-reactance circles with the |Γ| = 1 circle are the
purely imaginary impedance (0 + jX) points. (3) The center of the chart corresponds
to R = 1 point. (4) The perimeter of the chart corresponds to R = 0 points. (5) The
extreme right-hand point on the x-axis that is Γ = 1 corresponds to an open circuit.
(6) The extreme left-hand point on the x-axis that is Γ = −1 corresponds to a short
circuit. (7) Load impedances having a negative real part would be projected out
of the chart, that is, out of the |Γ| = 1 circle. Evidently for these impedances the
corresponding reflection coefficient would have an amplitude greater than unity.
Example 7.9 Show Z1 = 50+ j50 and Z2 = 50− j50 on the Smith chart. Assume
the line impedance is equal to 50 Ω.
Solution:
We begin by calculating the normalized values of the impedances,
Z1
Z1n = = 1+ j (7.159)
50
7.10 Smith Chart Mapping 325
and
Z2
Z2n = = 1− j (7.160)
50
which can be found using the Smith chart by finding the intersections of R = 1 and
X = 1, and R = 1 and X = −1 circles, respectively. This procedure is shown in
Figure 7.51.
Γi
+j 1
1+j
0 0.2 0.5 1 2 5 Γr
1-j
-j 1
Recalling Equation 7.72, one can write the impedance value at a distance d from the
load as follows
!
1 + ΓL e−2jβ `
1 + Γ(−d)
Z(−d) = Z0 = Z0 (7.161)
1 − Γ(−d) 1 − ΓL e−2jβ `
As it is evident from Equation 7.161, if one is moving toward the generator, by rotating
an amount of 2β ` clockwise from the initial impedance point, he/she would arrive at
the new impedance point on the chart. Otherwise, if one is moving from the initial
point toward the load, by rotating an amount of 2β ` counterclockwise, he/she would
arrive at the new impedance point on the chart.
326 Chapter 7. Transmission Lines and Impedance Matching
Example 7.10 Consider the normalized impedance Z = 0.5 + j0.7 and assume
that a −1 j reactance is added in series to this impedance. Sketch the initial and the
resulting impedances on the Smith chart.
Solution:
The initial and the final points are shown on the Smith chart in Figure 7.52.
Now moving on the R = 0.5 circle from the initial X = j0.7 toward X = − j0.3,
we come from the initial point to the final point on the chart.
Γi
+j 1
0.5+j0.7
0 0.2 0.5 1 2 5 Γr
0.5-j0.3
-j 1
We now turn our attention to Admittance Smith chart, that is, constant-conductance
and constant-susceptance circles. Recall that
1
Yn −1 1 −Yn
Γ= 1
= (7.163)
Yn +1 1 +Yn
If we imagine having a load whose normalized impedance is inverse of the given
normalized admittance (Zn = Y1n ), the Admittance Smith chart could be obtained by the
same procedure as the Impedance Smith charts just by replacing Γ by −Γ as shown in
Equation 7.164. In other words, if we rotate the impedance Smith chart by 180◦ , the
same chart would represent the Admittance Smith chart.
1
Yn −1 1 −Yn
ΓZ = 1
= = −Γ (7.164)
Yn +1 1 +Yn
7.10 Smith Chart Mapping 327
As such, the conductance and the susceptance circles can be derived from the resistance
and the reactance circles, respectively. This procedure is useful in the sense that once
we desire to add any reactance or resistance in series, we use the Impedance Smith
chart and once we desire to add any susceptance or conductance in parallel, we use the
Admittance Smith chart. A complete Smith chart is shown in Figure 7.53, in which the
blue and the red circles denote the impedance and admittance circles, respectively.
Care should be taken by the reader while using the Smith chart. Notice that wherever
on the chart the reactance is positive (upper half of the chart), then the susceptance is
negative and wherever on the chart the reactance is negative (lower half of the chart),
then the susceptance is positive. Furthermore, wherever on the chart the normalized
resistance is greater than unity, then the normalized conductance is less than unity
and wherever the normalized conductance is greater than unity, then the normalized
resistance is less than unity.
7.10.1 Some simple application rules while using the Smith chart
Rule 1: For matching, we generally use circular paths on the chart.
Rule 2: Rotating clockwise on a constant resistance circle is equivalent to adding a
series inductance.
Rule 3: Rotating counterclockwise on a constant resistance circle is equivalent to
adding a series capacitance.
Rule 4: Rotating clockwise on a constant conductance circle is equivalent to adding a
shunt capacitance.
Rule 5: Rotating counterclockwise on a constant conductance circle is equivalent to
adding a shunt inductance.
Rule 6: Rotating clockwise on a constant radius circle about the chart center is
equivalent to approaching the generator on the transmission line.
Rule 7: Rotating counterclockwise on a constant radius circle about the chart center is
equivalent to approaching the load on the transmission line.
Rule 8: Jumping from a constant resistance circle to another constant resistance
circle (on the real axis) is equivalent to putting a transformer on the transmission line,
meaning that increasing or decreasing the input resistance (depending on the sense of
the jump).
Rule 9: Jumping from a constant conductance circle to another constant conductance
circle (on the real axis) is equivalent to putting a transformer on the transmission line,
meaning that increasing or decreasing the input conductance (depending on the sense
of the jump).
Rule 10: Just rotating by 180◦ , one can read the admittance value on the same chart
(once he/she has recorded the impedance value on the impedance chart). Otherwise,
just rotating by 180◦ , one can read the impedance value on the same chart (once he/she
has recorded the admittance value on the admittance chart).
Rule 11: Using quarter wave transformers, one can travel clockwise on a half-circle
from the starting point at the load impedance to the desired input impedance at the end
point. As seen in Figure 7.54, this way one can transform a high impedance (point A)
to a low impedance (point B), or one can transform a low impedance (point C) to a
high impedance (point D).
The application of these rules is illustrated in Figure 7.54. We now revisit
example 7.11 and repeat it using the Smith chart.
328 Chapter 7. Transmission Lines and Impedance Matching
Γi
+j 1
-j 1
0.2 0.2
0.5 0.5
1 1
2 2
5 5
Γr
+j 1
-j 1
Example 7.11 Consider the circuit depicted in Figure 7.34. Use the Smith
chart to determine a matching network at 75 MHz which incorporates DC current
blocking.
Solution:
The load impedance consisting of a parallel resistor and a capacitor is depicted on
the admittance chart as YL = 1.7 + j18.9 mS or in normalized form ȲL = 0.085 +
j0.945 or its normalized impedance is depicted as Z̄L = 0.095 − j1.05 at point 1.
Now, we move on the constant normalized conductance circle of 0.085, on the
admittance chart such that we come across the R = 1 circle at point 2. We now
record the difference in normalized susceptances of point 2 and point 1 as
The inductive susceptance to be added in parallel with the load would become
then
% $
& '
1 1
L= = = 87.7 nH (7.167)
0.0242ω 0.0242 × 2π 75 × 106
1 1
C= = = 12.5 pF (7.168)
170ω 170 × 2π 75 × 106
With this series capacitance, we come to the center of the chart at point 3 as
depicted in Figure 7.55.
330 Chapter 7. Transmission Lines and Impedance Matching
Γi
2
Γr
3
1
1
Yn1=0.085+j0.945 Zn1=0.095-j1.05
Yn2=0.085-j0.265 Zn2=1+j3.4
Yn3=1 Zn3=1
Example 7.12 Find a matching network to match the load impedance illustrated
in Figure 7.56 to 50 Ω at 2.5 GHz. The reference impedance is 50 Ω.
0DWFKLQJ
=LQ =/ &S S) 5S ȍ
QHWZRUN
Solution:
We first compute the normalized admittance at 2.5 GHz
Γi
Γr
1
0.25
Yn1=0.25+j0.785 Zn1=0.37-j1.16
/
=LQ & =/ &S S) 5S ȍ
0DWFKLQJ
QHWZRUN
Figure 7.59 explains how the proper value of the inductor can be found using
the Smith chart.
Γi
Γr
Yn1=0.25+j0.785 Zn1=0.37-j1.16
Yn2=1-j1.31 Zn2=0.37+j0.48
Γi
3
Γr
Yn1=0.25+j0.785 Zn1=0.37-j1.16
Yn2=1-j1.31 Zn2=0.37+j0.48
Yn3=1 Zn3=1
Figure 7.60: Adding a capacitor to the load to reach the point 3 from
point 2.
Now, we compute the required values of the series inductance and the parallel
capacitance as follows
Cω ȲCY0 1.31(0.02)
ȲC = ⇒C = = = 1.67 pF (7.171)
Y0 ω 2π × 2.5 × 109
The reader might notice that using the Smith chart he/she would need only simple
arithmetic calculations to design the matching circuit.
7.10 Smith Chart Mapping 333
Example 7.13 The structure depicted in Figure 7.61 can be shown to match 1 kΩ
to 50 Ω at 2 GHz. The relative permittivity is considered as unity. Verify this fact
using the Smith chart.
ȍ
ȍ
Figure 7.61: The matching network in Example 7.13.
Solution:
We first record the load impedance on the Smith chart for the first section of the
transmission line:
From point 0 on the Smith chart, we turn clockwise, for 0.24λ (or 173◦ ) to reach
the point 1 and we record the input impedance as 0.501 − j0.047, at point 1. For
moving along the 300 Ω line, we normalized this impedance by the new reference
impedance of 300 Ω as
500
Zn2 = (0.501 − j0.047) × = 0.835 − j0.0783 (7.173)
300
We report this value on the chart as point 2. Now we turn clockwise about the
center of the chart by 0.24λ (or 173◦ ) to reach the point 3. We record the value of
the impedance at this point as 1.167 + j0.135. For moving along the 115 Ω line,
we normalized this impedance by the new reference impedance of 115 Ω as
300
Zn4 = (1.167 + j0.135) × = 3.047 + j0.351 (7.174)
115
We report this value on the chart as point 4. Now we turn clockwise about the
center of the chart by 0.35λ (or 252◦ ) to reach the point 5 on the chart. We record
the value of the impedance at this point as 0.445 + j0.569. We denormalize the
value of the impedance here to achieve
(51.26 + j65.4)
Zn6 = = 1.024 + j1.308 (7.176)
50
This value is reported as point 6 on the chart. By adding a series capacitive reac-
tance of − j65.4 ( C11ω = 65.4), we approximately reach to the center of the chart at
point 7. The same procedure could be followed analytically as follows:
By adding the series 1.21 pF capacitor, the final input impedance becomes
Γi
5 6
1 0 4
Γr
2 73
Zn0=2
Zn1=0.501-j0.047
Zn2=0.835-j0.078
Zn3=1.167+j0.135
Zn4=3.047+j0.351
Zn5=0.445+j0.569
Zn6=1.024+j1.308
Zn7=1.024≈1
Figure 7.62: The steps of matching in the circuit shown in Figure 7.61.
ȍ ȍ
= = =
Figure 7.64 shows the contour of this impedance matching on the Smith chart.
īL
= = =
īU
Figure 7.64: The contours of impedance matching on the Smith chart using
three sections of quarter wave transformers.
īL
% $
īU
Figure 7.65: Half-circle contour showing the quarter wave impedance trans-
formation on the Smith chart.
īL
&RQVWDQW4
ORFXV
4
īU
√
2
Figure 7.66: The impedance transformation contours and Q = 2 locus on
the Smith chart.
338 Chapter 7. Transmission Lines and Impedance Matching
Example 7.15 The structure depicted in Figure 7.67 can be shown to match 1 kΩ
to 50 Ω.
50Ω C1
+ L 1000Ω
VS C2
- 50Ω
Then
C2
= 3.47 (7.188)
C1
and
C1 +C2
L= = 4.58 nH (7.190)
C1C2 ω 2
7.11 Conclusion 339
(b) Given the quality factors of the inductor and the capacitors, the parallel equiva-
lent resistance is equal to (verification of the following is left to the reader):
"
# " #
C1 2 C2 2
QC QC
RP = RL k (QL Lω) k 1+ k RS k 1+ = 331.9 Ω
C1 ω C2 C2 ω C1
(7.191)
ω0 ω0 Lω0 2
Q= ⇒ BW = = = 86.7 MHz (7.192)
2πBW 2πQ 2πRP
The following figure shows the frequency response and the bandwidth of this
matching circuit through ADS simulation.
4
Vout/VS(dB)
3
2
1 BW=86.7MHz
0
-1
1
0.95
0.96
0.97
0.98
0.99
1.01
1.02
1.03
1.04
1.05
Frequency (GHz)
Figure 7.68: The frequency response and the bandwidth of the matching
network (ADS simulation result).
7.11 Conclusion
Impedance matching is crucial in RF circuits, to transfer the maximum power from
the source to the load. The impedance matching can be achieved through the use of
reactive elements such as inductors, capacitors, or transformers. Transmission lines are
widely used in RF circuits as well. Using the transmission line equations or the Smith
chart, we can use them for impedance matching of the RF circuits as well. Note that
while using the Smith chart in the design of the matching networks, generally inductors
and capacitors are considered to be lossless to have a quick insight and a quick design
340 Chapter 7. Transmission Lines and Impedance Matching
procedure. Care should be taken while using the Smith chart not to approach the
open-circuit vicinity on the impedance chart and not to approach the short-circuit
vicinity on the admittance chart, in order to avoid significant errors in the matching
network designs or evaluations. Open-circuit stubs or short-circuit stubs can be used
as the reactive elements in the matching circuits. Quarter-wave transmission lines
can be used as impedance transformers in the matching circuits as well. Furthermore,
the combination of the lumped elements and transmission lines could be used in the
matching circuit networks in any case.
7.13 Problems
Problem 7.1 Assume that a transmission line with a characteristic impedance of Z0
is terminated to 75 Ω. If SW R = 1.5, determine the characteristic impedance, Z0 .
Problem 7.2 Consider the circuit shown in Figure 7.69. Find the reflection coeffi-
cient, the input impedance Zin (λ /8), and SWR. Also, find the voltage and the current
phasors at the source and the load sides. What is the delivered power to the load?
100Ω Zin(λ/8)
+
10 0º Z0=50Ω ZL=100+j100Ω ZL
-
d=λ/8 d=0
Figure 7.69: A transmission line with the source and the load terminations.
Problem 7.3 Consider the amplifier shown in Fig. 7.70 with the input and output
matching networks. If the input impedance can be modeled as 15 KΩ k 15 pF and the
output equivalent circuit can be modeled as 410 Ω k 5 pF and the operation frequency
is 50 MHz, calculate the bandwidth at the input and at the output. Also calculate the
input matching circuit loss in dB at the center frequency. All capacitors have a quality
of factor of 100 at 50 MHz. Hint: The power delivered to the amplifier is the same
power delivered to the 15 kΩ.
C1 = 13 pF,C2 = 34 pF,C3 = C4 = 28 pF
L1 = 436 nH(Q = 11), L2 = 336 nH(Q = 32)
=LQ =RXW
Problem 7.4 Consider the load impedance at the point Z = (30+ j40) Ω on the Smith
chart and also the two matching networks shown in Figure 7.71. Plot the matching
circle paths on the Smith chart, and estimate the input impedance, and then compare
it to your analytical computations. The operating frequency is 1 GHz. Is there any
deficiency in these matching networks?
342 Chapter 7. Transmission Lines and Impedance Matching
5pF Z 6nH Z
Zin Zin
C1 C2
C2
L Z L Z
C1
(a) (b)
Figure 7.72: Two possible matching circuits for the specified load.
Problem 7.6 In the circuit shown in Figure 7.73, we desire to perform the matching
to the complex source impedance. Calculate the length of the line, `1 , the λ /8 stub
characteristic impedance, Z0,2 , and the type of the stub (open-circuit stub or short
circuit stub) for this purpose. The operation frequency is 1 GHz, and the effective
permittivity of the line is 3.
ȍ Mȍ
= ȍ
6WXE
96 Ȝ =/ M
= Ὓ
0DWFKLQJ
SODQH
Figure 7.73: The λ /8 single stub matching for a complex source impedance.
7.13 Problems 343
%LDV
FLUFXLWU\
&
/
&HUDPLF
ILOWHU &
ȍ__I)
d0
Z0=50Ω ZL=35-j47.5
l
Zin
Problem 7.13 Determine the values of `A and `B in Figure 7.76 such that the input
is matched to 50 Ω. Consider all the T-lines are of 50 Ω characteristic impedance. Note
that X could be either a short circuit or an open circuit.
λ/8
Z0=50Ω
ZL=60+j80
lB
50Ω
X
lA
Problem 7.14 Assuming a lossless 50 Ω line and the load impedance in the circuit
depicted in Figure 7.77, calculate the power delivered to the load. Hint: calculate first
the available power of the source and then the input reflection coefficient.
Z0
100Ω
+V 50Ω ZL=100+j100
5
d=λ/8
Figure 7.77: Determination of the delivered power to the load through a T-line.
7.13 Problems 345
Problem 7.15 A lossless line is terminated to a resistive 100 Ω load. If the reflected
voltage at the load side is V − (z = 0) = j Volts and the total current passing through
the load is equal to I = j (0.04) Amperes, determine the characteristic line impedance
Z0 (Fig. 7.78).
I(Z)
+
V(Z) Z0 100Ω
-
l
Z=- l Z=0
Problem 7.17 In the lossless T-line depicted in Figure 7.79 which is terminated to
150 Ω, the voltage phasor at endpoint of the line is V (z = 0) = j3 V.
1. Determine the complex functions Γ (z) ,V + (z) ,V − (z) along the line.
2. Determine the complex functions Z (z) ,V (z) , I (z) along the line.
3. Determine the values of the functions found in parts (1) and (2) at z = −1 m.
+
V(z) Z0=50Ω 150Ω
-
z=-1m z=0
Figure 7.79: The circuit for determination of the incident and the reflected
voltage phasors.
Problem 7.18 In the circuit depicted in Figure 7.80, compute Z(−`) as a function
of the T-line’s length and plot the results in two separate curves expressing the real and
the imaginary parts. Also, calculate the reflection coefficient at the points shown in the
figure (at the input, this value can be expressed in terms of `, where 0 < ` < λ2 ).
346 Chapter 7. Transmission Lines and Impedance Matching
l λ/4
Z(-l )
Z0=50Ω Z0=75Ω
25Ω
Γ Γ Γ
Problem 7.19 In the circuit depicted in Figure 7.81, we wish to perform the impedance
matching between a 100 Ω source and a 1000 Ω load using an ideal capacitor and an
ideal inductor at 1 GHz. Determine the required values of L, C, and the corresponding
Q of the circuit. Sketch the frequency response (| VVout |) of the circuit as well.
in
0DWFKLQJ
FLUFXLW
ȍ 9RXW
/
YLQ & Nȍ
Problem 7.20 Calculate the values of inductors and capacitors to achieve impedance
matching between 100 Ω source and 1000 Ω load considering a load’s L-section Q
factor of 10 at 1 GHz.
0DWFKLQJFLUFXLW
ȍ / /
9RXW
& & ȍ
YLQ
5HT
" =3cm
VDC=1V +- 150ȍ
Problem 7.22 An air line is depicted in Figure 7.84 which is loaded by an RLC
circuit. Draw the input reflection coefficient about 5 GHz (e.g., from 4.95 GHz to
5.05 GHz) on the Smith chart. What would be the equivalent circuit of this network
then?
L
L=3.186nH
Z0=50
r=1 r
" =15mm C=318fF
C
īin
L L
7ZRSRUW
9 9
QHWZRUN
i1 g g12 v1
= 11 (8.1d)
v2 g21 g22 i2
v1 A B v2
= (8.1e)
i1 C D −i2
The five above parameters can be easily measured at low frequencies by open- or
short-circuiting the network. Example 8.1 provides a better insight into this issue.
Example 8.1 Depicted in Figure 8.2 is a resistive two-port network. Derive the
Z-parameters according to Equation 8.1a.
=$ =% =$ =%
L =& Y Y =& L
Solution:
It follows from Equation 8.1a that
v1
z11 = i =0 = ZA + ZC (8.2a)
i1 2
v1
z12 = i =0 = ZC (8.2b)
i2 1
v2
z21 = i =0 = ZC (8.2c)
i1 2
v2
z22 = i =0 = ZB + ZC (8.2d)
i2 1
8.1 Representation of Two-Port Networks 351
9L 9L
where Γin is the input reflection coefficient of the two-port. The standard representation
of Equation
8.4is given by
b1 S11 S12 a1
= (8.8)
b2 S21 S22 a2
which is widely referred to as the S-parameters matrix. As with the other matrix
representations of circuits, each entry carries an interpretation of its own. S11 is the
input reflection coefficient once the output is terminated by the reference impedance,
S22 is the output reflection coefficient once the input is terminated by the reference
impedance, S21 is the forward transmission coefficient once the output is terminated
by the reference impedance, and S12 is the reverse transmission coefficient once the
input is terminated by the reference impedance. Each of the entries of the [S] matrix
can be derived individually
as
Vr1 b1
S11 = = (8.9a)
Vi1 Vi2 =0 a1 a2 =0
Vr1 b1
S12 = = (8.9b)
Vi2 Vi1 =0 a2 a1 =0
Vr2 b2
S21 = = (8.9c)
Vi1 Vi2 =0 a1 a2 =0
Vr2 b2
S22 = = (8.9d)
Vi2 Vi1 =0 a2 a1 =0
Equations 8.9a–8.9d also explicitly indicate how each one of the parameters can be
measured. To measure S11 , for instance, one needs to measure the ratio of the reflected
wave of port 1 to the incident wave of port 1 while the other ports are terminated by
the reference impedance. A 5-port network is depicted in Figure 8.4. The scattering
parameters for this network can be written as
b1 S11 S12 S13 S14 S15 a1
b2 S21 S22 S23 S24 S25 a2
b3 = S31 S32 S33 S34 S35 a3 (8.10)
b S
41 S42 S43 S44 S45
a
4 4
b5 S51 S52 S53 S54 S55 a5
Port 1
Port 2 Port 5
Device
Port 3 Port 4
Solution:
Using Equation 8.9a–8.9d, it follows that
Vr1 Zin − Z0
S11 = =ρ = (8.11)
Vi1 Vr2 =0 Zin + Z0
Terminating the output by the reference impedance (50 Ω), the input impedance is
given by
Hence,
50 − 50
S11 = =0 (8.13)
50 + 50
suggesting that we have matching at the input. Due to symmetry, S22 will also
assume a value of zero. Given the matching at the output (note that the reflected
wave at the output is zero), it follows that
(Z2 + Z0 ) k Z3 Z0 Z0
Vt2 = V2 = V1 = Vo
(Z2 + Z0 ) k Z3 + Z1 Z2 + Z0 Z2 + Z0
41.44 50
= V1 = 0.707 V1
41.44 + 8.56 50 + 8.56
(8.14)
Using the symmetry of the circuit, S12 = S21 = 0.707. Thus, the [S] matrix is
given by
0 0.707
[S] = (8.15)
0.707 0
354 Chapter 8. Scattering Parameters
Example 8.3 Find the S-parameters for the circuit depicted in Figure 8.6 at
1 GHz.
Solution:
It follows from Figure 8.6 that
1
Zin = R|| + 50 = 75 − j25 (8.17)
Cs
hence,
Zin − Z0
S11 = = 0.277∠ − 33.7◦ (8.18)
Zin + Z0
Zin
50Ω 50Ω 50Ω
+
+
Vs1 3.2pF Vo2
(a)
Zout
50Ω 50Ω 50Ω
+
+
Vo1 3.2pF Vs2
(b)
yielding
Zout − Z0
S22 = = 0.624∠ − 97◦ (8.20)
Zout + Z0
The forward gain and the reverse gain of the circuit can also be computed in a
similar manner as
1
VO2 2 Cs ||R
S21 = = 1 = 0.554∠ − 33.7◦ (8.21)
VS1 /2 Cs ||R + 2R
and
2 1 ||2R
VO1 R
S12 = = 1 Cs = 0.554∠ − 33.7◦ (8.22)
VS2 /2 Cs ||2R + R R + R
Example 8.4 Assume that in the oscillator depicted in Figure 8.7, an inductor
is tied to the base of the bipolar transistor, Q. The S-parameters of the two-port at
1 GHz have been measured as S11 = 1.8∠100◦ , S21 = 2.2∠−140◦ , S12 = 0.7∠140◦ ,
and S22 = 1.1∠ − 100◦ .
(a) Determine the maximum value for Rs so that the oscillation occurs.
(b) If L is changed in such a way that the input reflection coefficient becomes
Γin = −2, recalculate the maximum value of Rs .
ℓ=0.082λ
RS Z0=50Ω
L R=50Ω C=3.8pF
Γin ZL
Solution:
(a) Using Equation 7.72, the load impedance can be calculated as
1
R|| jCω + jZ0 tan (β `)
ZL = Z0 = 16.11 –j0.0496 (8.23)
1
Z0 + j R|| jCω tan (β `)
or
a2
Γin = S11 + S12 (8.26)
a1
a2
= S21 a1 + S22 a2 (8.28)
ΓL
Therefore,
a2 S21
= 1
(8.29)
a1 ΓL − S22
a2
Replacing a1 in Equation 8.26, the input reflection coefficient can be computed as
S21 S12 ΓL
Γin = S11 + (8.30)
1 − S22 ΓL
We can then deduce that Rs should be less than 15.83 Ω for the oscillation to occur.
8.2 Measuring S-Parameters Using a Network Analyzer 357
(b) The corresponding input impedance should first be calculated from the
given reflection coefficient as
For the oscillation to start, Rs should be less than the absolute value of the negative
resistance, i.e.,
1
RS < Z0 = 16.67 Ω (8.33)
3
6LJQDO
VRXUFH
6ZLWFKLQJ
FLUFXLW
'HW
'HW
$
$
G% G%
'HW 'HW
% %
G% G%
G% G%
4XLHVFHQW 4XLHVFHQW
ELDV ELDV
YROWDJH YROWDJH
3RUW 3RUW
'87
Therefore, all the scattering parameters can be evaluated using a network analyzer.
The measurements, however, are subject to errors which may stem from numerous
sources, as discussed earlier. In the next section, we discuss techniques to mitigate
these errors.
5HIHUHQFH 5HIHUHQFH
SODQH SODQH
'87
Ὓ Ὓ
D
7KURXJKOLQHIRU6PHDVXUHPHQW
Ὓ Ὓ
E
2SHQ
FLUFXLW
Ὓ Ὓ
F
6KRUW
FLUFXLW
Ὓ Ὓ
G
7HUPLQDWHG
Ὓ Ὓ
ȍ ȍ
Figure 8.9: Calibration circuits needed in the calibration process for a network
analyzer to measure the two-port S-parameters of the DUT.
8.2.4 One-Port and Two-Port Calibration for Short Circuit, Open Circuit, and
the Characteristic Impedance
Calibration is of great importance if a network analyzer is to accurately measure an
unknown impedance. This can be achieved by using components which simulate the
behavior of an open circuit, a short circuit, or reference impedance. After calibration,
only a slight difference can be observed between the real and the measured parameters.
If carried out for only one port, the above procedure can only provide the input or the
output matching accurately. To measure the device’s gain or the isolation, however, the
procedure should be performed for both ports. The calibration procedure is illustrated
in Figure 8.9.
Consider we intend to measure the S-parameters of a transistor as depicted in
Figure 8.9(a) where the transistor is accessed through the ports of the network analyzer
with two transmission lines with the lengths `1 and `2 , respectively.
As in Figure 8.9(b) a through-line with a length of `1 +`2 is realized, and it is referenced
as S21 = 1∠0, in order to measure the parameters S12 and S21 with a correct phase.
As in Figure 8.9(c), two open circuits are realized at the ports 1 and 2 of the network
analyzer through two transmission lines with the lengths `1 and `2 , respectively. As
such, here, at port 1, the reflection coefficient is referenced as S11 = 1∠0, and at port 2,
the reflection coefficient is referenced as S22 = 1∠0.
As in Figure 8.9(d), two short circuits are realized at ports 1 and 2 of the network
analyzer through two transmission lines with the lengths `1 and `2 , respectively. As
such, here, at port 1, the reflection coefficient is referenced as S11 = 1∠180◦ , and at
360 Chapter 8. Scattering Parameters
Solution:
(a) The error is equal to the delay generated by a reference transmission line with
the length of λ /8, that is about 90◦ , in reflection (assuming a lossless T-line).
(b) For proper measurement, the delay resulting from the transmission line should
be deduced from the phase of S11 . In other words
0
S11 = S11 e−j2β l (8.34)
Ɛ Ȝ
= ȍ
& &
5HIHUHQFH 5HIHUHQFH
SODQH SODQH
6
6’
2SHQFLUFXLW
0HDVXUHPHQW
FDOLEUDWHGDW&
UHIHUHQFHSODQH
Figure 8.11: The proposed model for the Device Under Test.
0
where S11 is the measured reflection coefficient at C1 reference plane, and S11
is the measured reflection coefficient at C2 reference plane.
0
S11 = S11 ej2β l (8.35)
In other words, the measured plots of S11 at C1 plane should be rotated 90◦ counter
clockwise to obtain the value of S11 at C2 reference plane.
Example 8.6 The S-parameters (S11 and S22 ) for the circuit shown in Figure 8.11
in the frequency range of 500 kHz–3 GHz have been measured as
Sketch a reasonable and proper equivalent circuit for this network. The boxes
contain passive devices.
Solution:
For an RF transistor, the input and the output impedances at 500 kHz are almost the
same as the DC impedances. Given the fact that S11 is real and is equal to −0.111,
then it could be deduced that a real impedance of the following value is seen at the
input at low frequencies:
1 + S11
Zin = Z0 = 40 Ω (8.40)
1 − S11
As the input impedance on the chart progressively becomes capacitive and finally be-
comes inductive, we suggest the input equivalent circuit as depicted in
Figure 8.13. This means rπ = 40 Ω.
362 Chapter 8. Scattering Parameters
īL
%
$ &
īU
'
1 + S11
Zin = Z0 ≈ 20 + j8.27 (8.41)
1 − S11
Then
Cπ ≈ 1.33 pF (8.43)
jrπ2 Cπ ω
Xin = jLb ω − ≈ j8.27 (8.44)
1 + (ωrπ Cπ )2
Then
Lb ≈ 1.5 nH (8.45)
8.2 Measuring S-Parameters Using a Network Analyzer 363
Lb Lc
+
Cπ rπ Vx gmVx ro Co
Figure 8.13: The proposed equivalent circuit for the measured S-parameters
of the transistor.
By the fact that S22 (the output impedance) is real at 500 kHz, again one could
say that the output impedance at DC is almost
1 + S22
Zout = Z0 = 500 Ω (8.46)
1 − S22
It is seen that S22 becomes gradually capacitive and its amplitude is slightly reduced,
so we suggest the output equivalent circuit as depicted in Figure 8.13. This means
ro = 500 Ω.
At 3 GHz, the output impedance becomes
1 + S22
Zout = Z0 ≈ 250 − j231.2 (8.47)
1 − S22
Then
Co ≈ 0.106 pF (8.49)
jro2Co ω
jLc ω − ≈ − j231.2 (8.50)
1 + (ωroCo )2
Then
Lc ≈ 1 nH (8.51)
364 Chapter 8. Scattering Parameters
9E
9E
5)&
5)&
Ὓ Ȝ
Ὓ Ȝ 3RUW
3RUW ȍ
ȍ
Figure 8.14: The effect of transmission line on the S-parameters.
Example 8.7 Derive the S-parameters for the transistor shown in Figure 8.14.
The S-parameters measured at ports one and two are given at 1 GHz as S11 =
0.3∠ − 20◦ , S21 = 2∠20◦ , S12 = 0.02∠ − 20◦ , and S22 = 0.15∠ − 62◦ . The T-lines
are lossless.
Solution:
The new S11 is given by
0 4π λ
S11 = S11 ej2β l1 = S11 ej λ 10 = 0.3∠ (−20◦ + 72◦ ) = 0.3∠52◦ (8.52)
It should be noted that forward gain and reverse isolation will also experience a
81◦ phase shift, yielding
0 2π λ
S12 = S12 ejβ (l1 +l2 ) = S12 ej λ ( 10 + 8 ) = 0.02∠ (−20◦ + 36◦ + 45◦ )
λ
= 0.02∠61◦ (8.54)
0 j 2π
λ ( 10 8 )
λ +λ
S21 = S21 e jβ (l1 +l2 )
= S21 e = 2∠ (20 + 36 + 45 ) = 2∠101◦
◦ ◦ ◦
(8.55)
Example 8.8 For the transistor model depicted in Figure 8.15, plot S11 in the fre-
quency range of 0 − 1 GHz on a polar coordinate, in 50 Ω reference impedance sys-
tem. Put a cross mark on the frequencies 50 MHz, 200 MHz, 500 MHz, 850 MHz,
1000 MHz.
8.2 Measuring S-Parameters Using a Network Analyzer 365
Z2
i1 Z1 2.4pF Z3
i2 i2-3Vx
+ Cx +
C1 C2
Port 1 400fF 75Ω Vx V2 -3Vx 100Ω 200fF Port 2 Z0
- -
Zin
Solution:
To compute S11 , we should first calculate the input impedance while the output is
loaded by Z0 . It can be found from Figure 8.15 that
Vx
i1 = + i2 (8.56)
Z1
or
Vx Vx −V2
i1 = + (8.57)
Z1 Z2
Having
or
Vx 1
Zin = = 1
(8.60)
i1 1 Z2 −3
Z1 + Z12 − Z
1+ Z 2
L
īL
īU
Figure 8.16: The plot of S11 of the transistor model from 50 MHz to
1000 MHz.
where
1 1 1
Z1 = ||75 Ω, Z2 = , Z3 = ||100 Ω (8.62)
jω (400 fF) jω (2.4 pF) jω (200 fF)
Example 8.9 For a transistor, the output circuit model is shown in Figure 8.17.
Compute S22 in the frequency span of 100 MHz − 1.5 GHz.
Solution:
S22 is given by
1
− Z0 1 − Z0 jco ω + r1o
jco ω+ r1o
S22 = 1
= (8.63)
+ Z0 1 + Z0 jco ω + r1o
jco ω+ r1o
8.2 Measuring S-Parameters Using a Network Analyzer 367
ro co
200Ω 500fF
Zout
-4.00 0
(S22)(Deg)
|S22|(dB)
-4.25 -10
-4.50 -20
-4.75 -30
-5.00 -40
0.1 0.25 0.5 0.75 1 1.25 1.5
Frequency (GHz)
S22 phase and amplitude behavior is depicted in Figure 8.18 from the above
computation.
Example 8.10 Two lossless transmission lines have been used to connect two
ports of a transistor on a PCB to two SMA connectors for S-parameter measure-
ments (consider the transmission lines as airlines). If the reference plane in test 1
has been calibrated for S11 and S22 , and assuming Zout = 25 − j50 and Zin = 10 at
500 MHz, determine the real input and output impedance of the transistor on the
Smith chart. (You may use the Smith software to do this.)
Solution:
We first derive the S-parameters considering the reference plane for test 1 as
Zin − Z0 10 − 50 2
S11 = = = − = 0.67∠180◦ (8.64)
Zin + Z0 10 + 50 3
Zout − Z0 25 − j50 − 50 1 − j8
S22 = = = = 0.62∠−82.9◦
Zout + Z0 25 − j50 + 50 13
FP FP
60$ 60$
ȍ ȍ
īL
īU
Figure 8.20: The input reflection coefficient, depicted on the Smith chart
0
[(1): S11 , (2): S11 ].
and
0
S22 = S22 ej2β l = 0.62∠−34.9◦ (8.66)
0 0
By solving for Zin , we arrive at Zin = 11.89 − j21.19 and the reflection coefficient
becomes
0
0 Zin − Z0
S11 = 0 = 0.67∠228◦ (8.68)
Zin + Z0
8.3 Conversion of Network Matrices 369
īL
īU
Figure 8.21: The output reflection coefficient, depicted on the Smith chart
0
[(1): S22 , (2): S22 ].
The same steps should be taken for the output impedance. Hence,
0 0
! !
Zout + jZ0 tan (β `) Zout + j50 tan 2π
15
Zout = Z0 0 = 50 0 = 25− j50 (8.69)
Z0 + jZout tan (β `) 50 + jZout tan 2π
15
0
and solving for the output impedance, it yields Zout = 83.77 − j96.6 and again the
output reflection coefficient can be computed as
0
0 Z − Z0
S22 = out
0 = 0.62∠−34.9◦ (8.70)
Zout + Z0
S Z Y
(Z11 −Z0 )(Z22 +Z0 )−Z12 Z21 (Y0 −Y11 )(Y0 +Y22 )+Y12Y21
S11 S11 ∆Z ∆Y
2Z12 Z0 −2Y12Y0
S12 S12 ∆Z ∆Y
2Z21 Z0 −2Y21Y0
S21 S21 ∆Z ∆Y
(Z11 +Z0 )(Z22 −Z0 )−Z12 Z21 (Y0 +Y11 )(Y0 −Y22 )+Y12Y21
S22 S22 ∆Z ∆Y
8.4 Conclusion
Scattering parameters are mainly used for high-frequency measurement of active and
passive devices. These parameters are measured using the network analyzer. Network
analyzers can distinguish between the incident and the reflected waves at the two
ports of the Device Under the Test (DUT). The scattering parameters are determined
with respect to a reference impedance. This reference impedance in most of the
measurement equipment is chosen to be 50 Ω. The input and the output reflection
coefficients of a device, i.e., S11 and S22 can normally be shown on the Smith chart,
while the forward gain S21 and the reverse gain S12 of a two-port are depicted on polar
or Cartesian coordinates. It is convenient to transform the scattering parameters into
impedance/admittance matrices and vice versa. An RF transistor S-parameters can be
measured, using a network analyzer, through a calibration procedure where the effects
of the interconnects can be canceled out.
8.5 References and Further Reading 371
8.6 Problems
Problem 8.1 Determine the S-parameters for the two-port network depicted in
Figure 8.22.
100Ω
50Ω 50Ω
λ/8 λ/8
Port 1 Port 2
Problem 8.2 The input of a transistor is modeled as in Figure 8.23. Sketch S11 in a
50 Ω system on the Smith chart from 0 to 1 GHz.
L=34.5nH
=/ &) ȍ
/
Y6 &
Q+
9E
1. Find the proper values of CF and Lb in the model to give in the stated value of
S11 .
2. Determine C1 and C2 for the required matching.
3. If the input inductor loss is modeled by a series resistance of 5 Ω, and assum-
ing QC1 = QC2 = 100, estimate the bandwidth of the matching network and
calculate the matching circuit loss in this case, in dB.
8.6 Problems 373
(a) ( b)
Problem 8.5 The input of a transistor is modeled as either of the networks depicted
in Figure 8.26, plot S11 in a 75 Ω reference system on the Smith chart for each of the
three cases shown in the figure for the frequency range of 0 − 1 GHz (with steps of
100 MHz). Compare these three cases.
2.1nH
Problem 8.6 Considering the bipolar transistor model given in Figure 8.27, determine
the corresponding expressions for the S-parameters, in terms of the circuit parameters,
with a reference impedance of Z0 . First assume Cµ k rµ is negligible, and then repeat
the problem while taking into consideration their effect.
rμ
rbb’ cμ
B C
+
rπ cπ Vb’e gmVb’e ro
-
Problem 8.7 Compute the S-parameters of the network depicted in Figure 8.28 at
2 GHz.
1.59pF 1.59pF
50Ω 50Ω
Problem 8.8 Find an expression for the two-port S-parameters for a reference
impedance of Z0 as shown in Figure 8.29. For which values of L and C, S11 and
S22 will tend to zero?
Z0 L Z0
Port1 C C Port2
3XUH
PP WUDQVLVWRU PP
6
Z0
Z0
Problem 8.11 A pair of similar amplifiers are cascaded by two three dB 90◦ couplers
with the specified S-parameters as in Figure 8.32. Determine the overall S-parameters
between port 1 and port 2. What is the advantage of this architecture?
AMP
3dB 3dB
coupler AMP coupler
0 2 2 0 0 2 2 0
j j
2 2 2 2
2 0 0 2 SA,11 SA,12 2 0 0 2
j j
2 2 2 2
S C= S A= S C=
2 2 2 2
j 0 0 SA,21 SA,22 j 0 0
2 2 2 2
0 2 2 0 0 2 2 0
j j
2 2 2 2
=LQīLQ =RXWīRXW
ħ6
,QSXW 2XWSXW
(V PDWFKLQJ PDWFKLQJ
ħ/
$PSOLILHU
=VīV =/ī/
Figure 9.1: Illustration of the general case of a microwave amplifier with input
and output matching circuits.
in communication systems, the RF power is the most scarce quantity, and it should
compete with noise, so approximately in all RF stages, the power gain is the quantity
to be optimized. Let us now introduce the most important property of a microwave
amplifier, that is, its power gain. Three separate definitions are considered for the
power gain as follows.
(1) Transducer Power Gain is the power delivered to the load, ZL , divided by the
available power of the source, PAVS , i.e.,
PL GL |VL |2 |VL |2
GT = = 1 2
= 4G G
S L (9.1)
PAVS
4GS |IS | |IS |2
ΓL = Γout ∗ (9.2)
then the absorbed power of the load becomes equal to the available power of the output
PL = PAVO (9.3)
and if
Γin = Γs ∗ (9.4)
then the available power of the source will be absorbed by the input, that is,
In this case, the transducer gain will attain its maximum value. This is the simultaneous
conjugate match condition.
9.2 Specification of Amplifiers 379
In terms of device’s S-parameters and the source and the load reflection coefficients,
the transducer power gain can be expressed as
|S21 |2 1 − |ΓS |2 1 − |ΓL |2
GT = (9.6)
|1 − ΓS Γin |2 |1 − S22 ΓL |2
GT ≤ GA (9.9)
(3) Operational Power Gain is the ratio of the power absorbed by the load from the
network to the power delivered to the network from the source, which can be written
as
PL GL |VL |2 GL
GP = = = |AV |2 (9.10)
Pin Gin |Vin |2 Gin
Here, Gin is the input conductance, GL is the load conductance, VL is the load voltage,
Vin is the input voltage, and AV is the voltage gain. If ΓL = Γout ∗ , that is, PL = PAVO ,
then power gain will attain its maximum value. One can generally expect that
GT ≤ GP (9.11)
In case of simultaneous conjugate match, all these three quantities will converge to the
value, that is,
We shortly develop expressions for each one the above triple definitions in terms of
S-parameters.
380 Chapter 9. Amplifier Design Using S-parameters
9.3.1 Stability
Passive loads, in general, have a reflection coefficient with an absolute value less than
unity, whereas for active two-ports (such as biased transistors), S-parameters could be
such that the input and/or the output reflection coefficients can have an absolute value
larger than unity. In this case, the circuit can become potentially unstable. In other
words, this means the input resistance or the output resistance is negative, which can
lead to instability. That is |Γin | > 1, or |Γout | > 1. In case of a unilateral network, for
example, S12 = 0, the unstable case reduces to |S11 | > 1 and |S22 | > 1, each revealing
the presence of a negative input or output resistances.
Regarding the stability, one could say a two-port is unconditionally stable if the
input reflection coefficient’s absolute value is smaller than unity for all passive loads,
and the output reflection coefficient’s absolute value is smaller than unity for all passive
sources. Or in other words, if the real part of both the input and the output impedances
for all passive load impedances and for all passive source impedances are positive. We
should bear in mind that all the expressions used here are in the frequency domain,
and hence, they are valid only within a certain bandwidth. If a two-port network is
not unconditionally stable, there exists a combination of load and source impedances
that leads to a negative real part for the input or a negative real part for the output
impedances. In terms of the reflection coefficients the stability means, the input
reflection coefficient and the output reflection coefficient for passive loads/sources
should satisfy the following condition, i.e., for |Γs |<1 and |ΓL |<1, can be described as
S21 ΓL S12
|Γin | = S11 + ≤1 for all |ΓL | ≤ 1 (9.13)
1 − S22 ΓL
and
S21 Γs S12
|Γout | = S22 + ≤1 for all |ΓS | ≤ 1 (9.14)
1 − S11 Γs
Relations 9.13 and 9.14 suggest that if the load or the source terminations are
passive, the input and the output reflection coefficients should remain passive, that is,
with an absolute value less than unity. To check the stability condition, relations 9.13
and 9.14 can be solved by putting each of the left-hand sides of the inequalities equal to
unity. A two-port network is conditionally stable if there exists passive loads for which
the absolute value of the input reflection coefficient is greater than unity (|Γin | > 1),
and there exists passive sources for which the absolute value of the output reflection
9.3 Performance Parameters of an Amplifier 381
coefficient is greater than unity (|Γout | > 1). As such the concept of stability using
input and output reflection coefficients, the problem of stability can be easily handled,
taking into consideration the locus for |Γin | = 1, once the load impedance varies, and
the locus for |Γout | = 1 once the source impedance varies, on the Smith chart. It can be
shown that both loci are described by circles given by the following relations (These
circles correspond to relations 9.13 and 9.14, respectively, once the equality holds).
The locus in the source plane can be described by
(S11 − ∆S22 ∗ )∗
Cs = (9.17)
|S11 |2 − |∆|2
and the locus in the load plane can be described by
(S22 − ∆S11 ∗ )∗
CL = (9.20)
|S22 |2 − |∆|2
In Equations 9.16 and 9.20, ∆ is the S-matrix determinant.
∆ = S11 S22 − S12 S21 (9.21)
With S-parameters known at a particular frequency, using Equations 9.16 and 9.17, the
circle corresponding to |Γout | = 1 can be found on the source impedance plane (Smith
chart) and the circle corresponding to |Γin | = 1 can be found on the load impedance
plane (Smith chart). These two circles are called the source stability circle and the load
stability circle, respectively. This means that these circles are the boundary between
the stable and the unstable region.
Therefore, either the inside or the outside of the stability circles corresponds to
the stable region, in other words the region where |Γout | < 1 or |Γin | < 1, respectively.
If the output is loaded by ΓL = 0, then the input reflection coefficient would become
Γin = S11 . The following cases could occur:
382 Chapter 9. Amplifier Design Using S-parameters
|Γin|=1
|Γin|=1 |Γin|=1
|Γin|<1
|Γin|<1
|Γin|<1 rL
rL rL CL
CL CL
|Γin|<1
|Γin|=1
|Γin|=1
|Γin|<1
|Γin|<1 rL
CL rL CL
|S11|<1 |S11|<1
(d) (e)
Figure 9.2: Input stability circles on the load plane Smith chart for five cases (the
shaded areas correspond to the stable region), (a) for |S11 | > 1 and conditional
stability, (b) for |S11 | < 1, the stability circle intersecting the chart while not
comprising the chart center, and consequently, conditional stability, (c) for
|S11 | < 1, the stability circle intersecting the chart while comprising the chart
center, and consequently, conditional stability, (d) for |S11 | < 1, the stability
circle does not intersect the chart, and consequently, unconditional stability, and
(e) for |S11 | < 1, the stability circle comprises the whole chart, and consequently,
unconditional stability.
1. If |S11 | > 1, then obviously there is a load (ΓL = 0) for which the two-port
becomes unstable, and therefore, the two-port can be conditionally stable
(Figure 9.2(a)).
2. If |S11 | < 1, and if the stability circle intersects the Smith chart, and the center
of the chart (ΓL = 0) is outside the stability circle, the points inside the chart
and outside the stability circle correspond to the stable region. Anyway, the
two-port is conditionally stable in this case (Figure 9.2(b)).
3. If |S11 | < 1, and if the stability circle intersects the Smith chart, and the center
of the chart (ΓL = 0) is inside the stability circle, the points inside the chart and
inside the stability circle correspond to the stable region. Anyway, the two-port
is conditionally stable in this case (Figure 9.2(c)).
4. If |S11 | < 1, and if the stability circle does not intersect the Smith chart, and
the center of the chart (ΓL = 0) is outside the stability circle, the points inside
9.3 Performance Parameters of an Amplifier 383
the chart and outside the stability circle correspond to the stable region. Conse-
quently, the two-port is unconditionally stable (Figure 9.2(d)).
5. If |S11 | < 1, and if the stability circle does not intersect the Smith chart but
the whole chart is inside the stability circle (evidently the center of the chart
is inside the stability circle as well), the points inside the chart and inside the
stability circle correspond to the stable region. Consequently, the two-port is
unconditionally stable (Figure 9.2(e)).
The same argument can be used for the source plane and S22 , as depicted in Figure 9.3.
This leads to an important conclusion: if |S11 | < 1 and |S22 | < 1, the network is
unconditionally stable if and only if the stability circles at the load plane and at
the source plane do not intersect the Smith chart. These points are demonstrated in
Figures 9.2 and 9.3.
|Γout|=1
|Γout|=1
|Γout|=1
rS
|Γout|<1
|Γout|<1
CS rS
CS
|Γout|<1 rS
CS
|Γout|=1 |Γout|=1
|Γout|<1 rS
|Γout|<1 rS
CS CS
|S22|<1 |S22|<1
(d) (e)
Figure 9.3: Output stability circles on the source plane Smith chart for five
cases (the shaded areas correspond to the stable region), (a) for |S22 | > 1 and
conditional stability, (b) for |S22 | < 1, the stability circle intersecting the chart
while not comprising the chart center, and consequently, conditional stability,
(c) for |S22 | < 1, the stability circle intersecting the chart while comprising
the chart center, and consequently, conditional stability, (d) for |S22 | < 1, the
stability circle does not intersect the chart and consequently, unconditional
stability, and (e) for |S22 | < 1, the stability circle comprises the whole chart,
and consequently, unconditional stability.
384 Chapter 9. Amplifier Design Using S-parameters
As observed in Figure 9.2(b), when |S11 | < 1, the shaded area includes the origin,
while in Figure 9.2(a), when |S11 | > 1, only a small area within the Smith chart is
covered by the stability circle. The same point is illustrated for the output reflection
coefficient in Figure 9.3. In Figure 9.3(b), when |S22 | < 1, the shaded area includes
the origin, while in Figure 9.3(a), when |S22 | > 1, only a small area within the Smith
chart is covered by the stability circle.
Now, the necessary and the sufficient conditions for unconditional stability of a
two-port are that the stability circles do not intersect the Smith chart at the source and
at the load planes. That is
Given |S11 | < 1 and |S22 | < 1, the above conditions can be translated into the following
pair of conditions:
and
In Equation 9.24, k is called Rollett’s stability factor. Equations 9.24 and 9.25 are the
necessary and the sufficient conditions for stability of the two-port, given |S11 | < 1 and
|S22 | < 1. Or, in other words, the conditions are k > 1 and |∆| < 1.
provided S12 6= 0 and k > 1. Note that for the case where S12 = 0 one can use the
unilateral transducer power gain as in Equation 9.29. Now, consider when we have no
constraint on the two-port and the two-port is unconditionally stable, for simultaneous
conjugate match, we should have
S12 S21 ΓL
Γin = S11 + = Γs ∗ (9.27)
1 − S22 ΓL
9.3 Performance Parameters of an Amplifier 385
and
S12 S21 Γs
Γout = S22 + = ΓL ∗ (9.28)
1 − S11 Γs
The simultaneous resolution of Equations 9.27 and 9.28 which results in maximum
APG or the maximum transducer gain is called the simultaneous complex conjugate
match (we discuss the solution of these equations in the following). Therefore, for
design purposes, we first choose a specific transistor with required power-frequency
capability. Then, using the load and the source, proper matching networks can be
designed. We should also note that there exists a relationship between the input and the
output reflection coefficients through S12 , further complicating impedance matching.
For the sake of simplicity, we first consider the case where S12 = 0, which is the
property of a unilateral transistor.
where we have
and
There exists a solution if and only if B1 2 − 4|C1 |2 ≥ 0 which means k > 1. In order to
determine the sign before the square root term in Equation 9.34, the solution which
has an absolute value less than unity would be acceptable. In other words, we look for
the sign of B1 , if it is positive, then the square root sign is negative and vice versa. The
source impedance can be found from
1 + Γs
Zs = Z0 (9.36)
1 − Γs
Therefore
q
B2 ± B2 2 − 4|C2 |2
ΓL,O = (9.38)
2C2
where
and
There exists a solution if and only if B2 2 − 4|C2 |2 ≥ 0 which means k > 1. In order to
determine the sign before the square root term in Equation 9.38, the solution which
has an absolute value less than unity would be acceptable. In other words, we look for
the sign of B2 , if it is positive, then the square root sign is negative and vice versa. The
load impedance can be found from
1 + ΓL
ZL = Z0 (9.41)
1 − ΓL
It should be noted that, if one of the source or load reflection coefficients are found, the
other one can be found from either through Equations 9.31 or 9.30.
9.3 Performance Parameters of an Amplifier 387
Example 9.1 The S-parameters at the given bias are measured for a bipolar
transistor at 200 MHz and at VCE = 10V, IC = 10 mA as S11 = 0.4∠162◦ , S12 =
0.04∠60◦ , S21 = 5.2∠63◦ , S22 = 0.35∠ − 39◦ . Assume that the amplifier is termi-
nated to a 50 Ω impedance at both input and the output. Perform complex conjugate
matching such that the maximum transducer power gain occurs.
Solution:
First, we examine the stability conditions using the provided S-parameters
from Equations 9.24 and 9.25 as
Now, we compute k as
As we have k > 1, |∆| < 1, |S11 | < 1, and |S22 | < 1, the two-port network is
unconditionally stable. The MAG can be calculated from section 9.3.2 as
It follows that
S21 p
MAG = 10 log + 10 log k − k2 − 1 (9.45)
S12
5.2 p
= 10 log + 10 log 1.74 − 1.742 − 1 = 16.1 dB
0.04
For example, if the desired maximum gain was larger than 16.1 dB, this transistor
could not be used. We can find the load reflection coefficient assuming the complex
conjugate; we have
and
which results in
q q
B2 ± B2 2 − 4 |C2 |2 0.958 − 0.9582 − 4(0.377)2
ΓL,O = = = 0.487∠+39◦
2C2 2 (0.377∠−39◦ )
(9.48)
S21 ΓL S12 ∗
Γs,O = S11 +
1 − S22 ΓL
5.2∠63◦ (0.487∠39◦ ) 0.04∠60◦ ∗
◦
= 0.4∠162 + (9.49)
1 − 0.35∠ − 39◦ (0.487∠39◦ )
= 0.522∠ − 162◦
Γi
ΓL,O
Γr
Γs,O
ΓL,O=0.487 +39º
Γs,O=0.522 -162º
Figure 9.4: Input and output reflection coefficients on the Smith chart.
9.3 Performance Parameters of an Amplifier 389
Γi
A
Γr
C
B
YnA=1
YnB=1+j1.45
YnC=2.62+j1.45 ZC=Zs=16-j7
ZnA=1
ZnB=0.32-j0.468
ZnC=0.32-j0.14
and
|Im{ZC } − Im{ZB }| 0.328 × 50
L= = = 13nH (9.51)
ω 2π 200 × 106
respectively.
The same procedure can be carried out in order to find the output matching
network. We first choose a series capacitor and move counterclockwise, on the
constant resistance contour, from the origin to point B in Figure 9.6, then by choos-
ing a parallel inductor, we move counterclockwise, on the constant conductance
contour, to reach ΓL,O point. These steps are depicted in Figure 9.6.
The inductor and capacitor values can be similarly found as
1 1
L= = = 50.3nH
ω |Im{YC } − Im{YB }| 2π 200 × 106 0.79 × 20 × 10−3
(9.52)
390 Chapter 9. Amplifier Design Using S-parameters
Γi
C
A Γr
B
YnA=1
YnB=0.38+j0.485
YnC=0.38-j0.305 ZC=ZL=79.5+j64
ZnA=1
ZnB=1-j1.27
ZnC=1.59+j1.28
Vcc
50.3nH
12.5pF
+ 50Ω
23pF
and
1 1
C= = = 12.5pF (9.53)
ω |Im{ZB } − Im{ZA }| 2π 200 × 106 1.23 × 50
We now proceed to study the design of two-port amplifiers for a specific gain.
9.4 Power Gain Contours 391
where
Gp 1 − |ΓL |2
gp = 2
= (9.55)
|S21 | 1 − |S11 |2 + |ΓL |2 |S22 |2 − ∆2 − 2 Re (ΓLC2 )
where C2 is given by Equation 9.40. In the above equations, Gp and gp are a function
of S-parameters and the load reflection coefficient. It can be shown that those values
of load reflection coefficient which yield a constant gp lie on a circle, which we refer
to as constant OPG contour from now on. The equation of this contour on the load
reflection coefficient plane is given by (9.56):
ΓL −Cp = rp (9.56)
where Cp is the center and rp is the radius of the circle which are given by
gC∗
Cp = p 2 (9.57)
1 + gp |S22 |2 − |∆|2
and
q
1 − 2k |S12 S21 | gp + |S12 S21 |2 gp 2
rp = (9.58)
2 2
1 + gp |S22 | − |∆|
Equation 9.57
suggests that the distance from the origin to the center of the circle is
equal to Cp and its angle can be found by the angle of C2 ∗ . The maximum OPG occurs
392 Chapter 9. Amplifier Design Using S-parameters
where rp is equal to zero. Now, if we equate rp to zero, it follows from Equation 9.58
that
g2 p,max |S12 S21 |2 − 2k |S12 S21 | gp,max + 1 = 0 (9.59)
where gp,max is the maximum value of gp . Solving for gp for the unconditionally stable
case, we obtain
1 p
gp,max = k − k2 − 1 (9.60)
|S12 S21 |
This is indeed the case where the conjugate matched condition occurs at the output.
Now, substituting Equation 9.60 into Equation 9.55, we have
|S21 | p
Gp,max = k − k2 − 1 (9.61)
|S12 |
Provided |S12 6= 0| and k > 1. The minimum value for gp is equal to zero, which
means Gp = 0 as well. We can observe from Equation 9.55 that Gp = 0, if only if
the magnitude of the load reflection coefficient is equal to unity (ΓL = 1). In other
words, OPG is equal to zero if all the power is reflected by the load. For a given power
gain, the load reflection coefficient can be deduced from constant OPG contours. The
maximum gain occurs when the load reflection coefficient lies at the distance |ΓL,O |
and where gp,max = Gp,max /|S21 |2 . The maximum output power occurs when we have
complex conjugate matching at the input, i.e., Γs =Γin ∗ . We can say equivalently that if
Γs = Γin ∗ , the input power is equal to the maximum available input power. Therefore,
in this case, the maximum transducer gain is equal to the maximum OPG. The constant
power contour can be drawn as follows: (1) For a given power gain, we draw the
desired circle from Equation 9.56; (2) We choose a desired load reflection coefficient
on the contour; (3) For the given load reflection coefficient, the maximum output power
is obtained when we have complex conjugate matching at the input, i.e., Γs = Γin ∗ .
The resulting source reflection coefficient gives an OPG equal to the transducer power
gain.
Example 9.2 The S-parameters at the given bias are measured for a bipo-
lar transistor at 250 MHz and at VCE = 5V, IC = 5 mA as S11 = 0.277∠ − 59◦ ,
S12 = 0.078∠93◦ , S21 = 1.92∠64◦ , S22 = 0.848∠ − 31◦ . Assume that the input
and the output terminations are Zs = 35 − j60 and ZL = 50 − j50, respectively.
Design a two-port amplifier such that the gain of 9 dB is achieved at 250 MHz.
where
Γi
rP
CP
C
Γr
A
B
YnA=0.5+j0.5
YnB=0.1+j0.3
YnC=0.1-j0.125 ZC=195+j244
ZnA=1-j1 Cp=0.712 +33.9º
ZnB=1-j3 rp=0.285
ZnC=3.9+j4.88
Figure 9.8: The load matching using a series capacitor and a parallel inductor
to achieve 9 dB gain.
The amplifier is unconditionally stable with k = 1.033 and ∆ < 1. It can be
shown that the maximum gain is equal to 12.9 dB. We have
(9.66)
and
q
1 − 2k |S12 S21 | gp + |S12 S21 |2 gp 2
rp = = 0.285 (9.67)
2 2
1 + gp |S22 | − |∆|
The constant gain contour is shown in Figure 9.8. In order to arrive at a point on
the contour from the starting point, A (ZnA = 1 − j1), we move counterclockwise
on a constant resistance circle (R = 50 Ω) with a series capacitor to point B and
394 Chapter 9. Amplifier Design Using S-parameters
1 1
C= = = 6.4pF (9.68)
ω |Im{ZB } − Im{ZA }| 2π 250 × 106 2 × 50
and
1 1
L= = = 75nH (9.69)
ω |Im{YC } − Im{YB }| 2π 250 × 106 0.425 × 0.02
For complex conjugate matching at the input, the chosen load reflection coefficient
0
is ΓL = 0.82∠14.2◦ . The input reflection coefficient then would be
0
!∗
0 S21 ΓL S12
Γs = S11 + 0
1 − S22 ΓL
(1.92∠64◦ ) (0.82∠14.2◦ ) (0.078∠93◦ ) ∗
= 0.277∠ − 59◦ + (9.70)
1 − (0.848∠ − 31◦ ) (0.82∠14.2◦ )
= 0.105∠160◦
Γi
C
D
Γr
A
YnA=0.36+j0.62
YnB=0.36+j1.24 B
YnC=1.22-j2.01 ZD=41+j3
YnD=1.22-j0.09
ZnA=0.7-j1.2
ZnB=0.22-j0.74
ZnC=0.22+j0.36
ZnC=0.82+j0.06
Vcc
75nH
6.4pF
35nH Vout
ZS
ZL
+
ES 7.9pF 24.4pF ΓL′ ΓL
Γs Γs′
Figure 9.10: The input and output matching networks for complex source
and load impedances.
and
|Im{ZC } − Im{ZB }| 1.1 × 50
L= = = 35nH (9.72)
ω 2π 250 × 106
and
|Im{YD } − Im{YC }| 1.92 × 0.02
C= = = 24.4pF (9.73)
ω 2π 250 × 106
While in Example 9.2 the load and the source impedances were not equal to 50 Ω,
normally this is not the case and every stage is normally matched to the reference
impedance, e.g., 50 Ω. In the next example, we take into consideration both the stability
and the power gain contours.
Example 9.3 S-parameters at 200 MHz and bias information for transistor 2N5179
are given as VCE = 6V , IC = 5 mA, S11 = 0.4∠280◦ , S12 = 0.048∠65◦ , S21 =
5.4∠103◦ , S22 = 0.78∠345◦ . Also assume 50 Ω terminations. Design the two-port
amplifier such that 12 dB power gain is achieved at 200 MHz.
396 Chapter 9. Amplifier Design Using S-parameters
Therefore, the two-port is conditionally stable (k < 1). Recalling Equation 9.35,
we now derive the input and the output stability circles as
The center and the radius of the locus of the unity output reflection coefficient
(input stability circle) can be found from Equation 9.17:
C1 ∗ 0.241∠136.6◦
Cs = 2 2
= = 10∠−43.4◦ (9.78)
|S11 | − |∆| 0.42 − 0.4292
The center and the radius of the locus of the unity input reflection coefficient (output
stability circle) can be found from Equation 9.20:
C2 ∗ 0.65∠24◦
CL = 2 2
= = 1.53∠24◦ (9.80)
|S22 | − |∆| 0.782 − 0.4292
We then proceed to find the center and the radius corresponding to the desired gain
in the example. It follows from Equation 9.55:
Gp 101.2 15.85
gp = 2
= = = 0.543 (9.82)
|S21 | 5.42 5.42
The areas of interest are shown in the Smith chart in Figure 9.11. We should choose
a point on the constant gain contour such that it lies outside the output stability
circle. Let’s choose ΓL = (0.724 − 0.287) ∠(24 + 180)◦ = 0.437∠204◦ to be sure
of the output stability. Now we choose a proper value for Γs such that the transducer
gain becomes equal to the operating power gain (the input is matched to the source),
so it can be calculated as follows
S12 S21 ΓL ∗
Γs = S11 + = 0.409∠68◦ (9.85)
1 − S22 ΓL
|īin|=1
īi īi Constant
Input gain
unstable circle
region Output
|īout|=1
unstable
region
rL
īs CL
CP
īr
rP
īr
īL
(a) (b)
Figure 9.11: (a) The input stability circle and the corresponding source
reflection coefficient. (b) The output stability circle and the chosen load
reflection coefficient.
398 Chapter 9. Amplifier Design Using S-parameters
GA 1 − |Γs |2
ga = = (9.87)
|S21 |2
1 − |S22 |2 + |Γs |2 |S11 |2 − ∆2 − 2real{ΓsC1 }
and
In a similar way to those expressions already derived earlier for OPG, the center and
the radius of constant APG circles which are drawn on the input reflection coefficient
plane can be, respectively, found as
gC∗
Ca = a 1 (9.89)
1 + ga |S11 |2 − |∆|2
and
q
1 − 2k |S12 S21 | ga + |S12 S21 |2 ga 2
rp = (9.90)
2 2
1 + ga |S11 | − |∆|
At the end, the circles can be plotted using Ca and rp on the input reflection coefficient
plane with every point on the circle yielding the desired gain. The maximum gain is
achieved when the load and the output reflection coefficient are conjugate matched, in
which case the transducer power gain (TPG) and APG become equal. Interestingly,
as constant noise contours are also drawn on the input reflection coefficient plane, a
compromise can be made between the available gain and the minimum noise figure on
this plane. As a result for the amplifiers where the operating power gain is critical, we
use the procedure described in Example 9.3, and for the amplifiers where the APG and
the noise figure are of more importance, we use the procedure described in the next
section.
9.5 Noise Behavior of a Two-Port Network 399
where k is the Boltzmann constant of a value of 1.374 × 10−23 ◦Jk , T is the absolute
temperature of the resistor in Kelvins, and B is the operating bandwidth. As suggested
by Equation 9.91, the noise power is a direct function of its bandwidth and the
bandwidth is normally limited value in an RF system. Thermal noise is widely known
XS XS en
*
+
-
RS Noisy RS Noiseless
two-port ZL in two-port ZL
+ network + network
Vn
* -
Vn
* -
2 2
Vn =4KTBRS Vn =4KTBRS
as a white noise, as it contains all frequency components and has a flat spectral behavior
over the frequency range. The available noise power of a resistor can be derived as
V 2 n,rms
PN = = KT B (9.92)
4RS
Note that noise voltage has a random value, its instantaneous value is not known, but
its rms value can be estimated. The wider the bandwidth, the narrower and larger the
instantaneous voltage spikes.
Example 9.4 Find the available noise power which a resistor generates at the
standard temperature, i.e., T0 = 290◦ k in a bandwidth of 1 Hz. Then, calculate the
noise voltage and power for a 2 MΩ resistor in a bandwidth of 5 kHz and at the
standard temperature.
Solution: We have from Equation 9.92,
PN = KT B = 1.374 × 10−23 (290) (1) = 3.985 × 10−21W
(9.93)
and finally, the noise power can be calculated from Equation 9.93 as
2
PN 12.6 × 10−6
PN (dBm) = 10 log −3 = 10 log × 103
10 4 2 × 106
= 10 log 19.9 × 10−15 = −137 dBm (9.96)
In Equation 9.98, PSO is the output signal power and PSi is the input signal power.
Therefore, Equation 9.97, can be rewritten as
PSi /PNi SNRi
F= = (9.99)
PSO /PNO SNRO
In other words, the noise figure can be defined as the ratio of the SNR at the input
to the SNR at the output. Given the fact that the noise figure is always greater than
unity, the output SNR is always smaller than the input SNR, so it is evident that in an
amplifier one cannot improve the SNR, as such an RF engineer would rather strive not
to degrade it. The curious student might ask why do we ever amplify the signal in a
receiver where at every stage the SNR would be degraded within the amplifier chain?
The reason is that to be able to further process the signal, it is necessary that the signal
should attain a certain required level (a few dBm’s, for example) for detection. In an
amplifier, in order to find the minimum noise figure, the choice of a proper source
reflection coefficient is important. In Figure 9.13, a model is provided for calculation
of noise figure in cascaded amplifiers. As depicted in Figure 9.13, PNi is the input
noise power, GA1 and GA2 are the APGs of the first and the second stage, and PN1 and
PN2 represent the available noise powers at the output of each amplifier which result
from its own intrinsic noise. Hence, the total noise at the output can be written as
PNO = GA2 GA1 PNi + PN1 + PN2 (9.100)
First Second
amplifier amplifier
R PNi=KTBR GA1PNi+PN1 PNo ZL
GA1 GA2
PN1 PN2
Figure 9.13: Model for noise figure calculation in a cascade of two stages.
402 Chapter 9. Amplifier Design Using S-parameters
and
PN2
F2 = 1 + (9.104)
PNi GA2
F1 and F2 are the noise figures corresponding to each one of the stages. Equation 9.102
suggests that the noise of the second stage is diminished by a factor of GA1 . We
then realize that the noise of the second stage cannot affect the overall noise figure
significantly. This leads to an important observation from the design point of view: the
noise of the first stage is the most important part of the overall noise figure, given the
fact that the power gain of the preceding amplifiers masks the noise of the following
stages, and therefore it is appropriate to diminish the noise figure of the first stage at
the price of losing a little bit of the power gain. In fact, we can target for the minimum
noise figure for a lesser power gain at the first stage. A trade-off always exists between
NF and the APG in any design. The design can be made such that the minimum NF
is obtained. Consider two amplifiers with NF’s and gains of F1 , F2 , GA1 , and GA2 ,
respectively. If the first amplifier precedes the second amplifier, the overall NF denoted
by F12 can be written as
F2 − 1
F12 = F1 + (9.105)
GA1
and if the second amplifier comes first, F21 is given by
F1 − 1
F21 = F2 + (9.106)
GA2
To have a lower NF in the first case, i.e., F12 < F21 , we should have
F2 − 1 F1 − 1
F1 + < F2 + (9.107)
GA1 GA2
which can be rewritten as
F1 − 1 F2 − 1
1
< (9.108)
1 − GA1 1 − G1A2
M1 < M2 (9.109)
F −1
M= (9.110)
1 − G1A
noise figure in this structure will be smaller. Hence, we predict that, in order to have
a minimal NF in a cascade of stages, the amplifier with the lower noise measure, M,
should precede the one with larger noise measure. It can be shown that NF for a
cascade of many stages is given by
F2 − 1 F3 − 1 F4 − 1
F = F1 + + + +··· (9.111)
GA1 GA1 GA2 GA1 GA2 GA3
which is called the “Friis’ relation” for the noise figure. This equation for the special
case of F1 = F2 = · · · Fn and GA1 = GA2 = · · · = GAn and, assuming an infinite chain
of identical amplifiers, reduces to
F1 − 1
F = 1+ = 1 + M1 (9.112)
1 − G1A1
This means the noise figure of an infinite chain of identical amplifiers tends to a limited
value (the noise measure plus unity).
Rn 2
F = Fmin + Ys −Yopt (9.113)
Gs
1 − Γs
YS = Y0 (9.114)
1 + Γs
and
1 − Γopt
Yopt = Y0 (9.115)
1 + Γopt
This relation is a function of Fmin , rn , and Γopt . These three parameters, two real and
one complex (or equivalently four real parameters), are widely known as the noise
parameters and are provided either in the datasheet by the vendor or can be found
404 Chapter 9. Amplifier Design Using S-parameters
by measurement. The input reflection coefficient can be altered and the noise figure
can be measured accordingly. Multiple measurements with different source reflection
coefficients (measured using a network analyzer), and the noise figure measured by a
noise figure meter will allow the engineer to extract the four required noise parameters.
Here Fmin is a function of the bias current or the bias voltage of the device and the
operating frequency. Therefore, only a single Γopt corresponds to every value of Fmin .
Equation 9.116 can be rewritten such that for a specific input reflection coefficient, ΓS ,
the resulting NF is, say F = Fi . Hence
Γs − Γopt 2
Fi − Fmin 2
2
= 1 + Γopt (9.117)
1 − |Γs | 4rn
As suggested by Equation 9.117, for a given noise figure, Fi , the right-hand side of the
equation is a constant. Therefore, if we define a factor, Ni as
Fi − Fmin 2
Ni = 1 + Γopt (9.118)
4rn
and, it follows that
Γs − Γopt 2
= Ni (9.119)
1 − |Γs |2
2 Γopt 2 Ni
|Γs |2 − Re Γs Γopt ∗ +
= (9.120)
1 + Ni 1 + Ni 1 + Ni
This represents a contour on the input reflection coefficient plane. Equation 9.120 can
also be rewritten alternatively as
2 N 2 + N 1 − Γ 2
Γ i i opt
Γs − opt =
(9.121)
1 + Ni (1 + N )2 i
This describes a circle on the input reflection coefficient plane with the center, CF,i ,
and the radius, rF,i , as follows
Γopt
CFi = (9.122)
1 + Ni
and
r
1 2
rFi = Ni 2 + Ni 1 − Γopt (9.123)
1 + Ni
Using Equation 9.118, Ni can be obtained for a given value of Fi . Ultimately, having
both the centers and the radii of constant noise figure contours, they can all be plotted
9.6 Constant Noise Figure Contours 405
Γi
3dB
Γopt A
3.5dB
4dB
5dB
6dB
Γr
Γopt=0.58 +138º
ΓA=0.38 +119º
on the input reflection coefficient plane. Equations 9.118 through 9.123 suggest that,
when Fi = Fmin , Ni is equal to zero, and the center corresponds to Γopt and the radius
would be zero (Fmin contour lies at the center point, Γopt , with a radius of zero). We
realize from Equation 9.120 that the centers of all other constant noise figure contours
lie on a line connecting the center of the chart to the Γopt point; in other words, they
lie on a line passing through the center with ∠Γopt . A typical group of constant noise
figure contours are depicted in Figure 9.14. As it is obvious from Figure 9.14, the
minimum noise figure, Fmin , corresponds to Γs = Γopt = 0.58∠138◦ and it is equal
to 3 dB. Any other neighboring point will have a higher noise figure. At point A, for
instance, with Γs = 0.38∠119◦ , we have NF = 4 dB.
In practical designs, there is always an unwanted discrepancy between the targeted
NF value and the NF derived from the measurement which stems from the matching
network imperfection and also inaccuracies in the transistors noise parameters’ mea-
surement. Typically speaking, this may amount to a few 0.1 dB’s to 1 dB change in
NF.
In a practical low-noise amplifier design, there exists a trade-off between the APG,
the noise figure, and VSWR, or equivalently, matching. The trade-off between the
noise and the power gain is illustrated in Figure 9.15, which the transistor of choice is
unilateral and a group of constant NF and constant gain contours are plotted. As it is
obvious from Figure 9.15, the maximum power gain and the minimum NF points do
not coincide in general. The normalized power gain in Figure 9.15 is Gs = 3 dB, which
occurs with Γs = 0.7∠110◦ and results in Fi = 4 dB. The minimum NF, Fmin = 0.8 dB
is achieved with Γs = 0.6∠40◦ , and corresponds to normalized Gs = −1 dB.
Using the above constant gain and the constant noise figure contours, one can
easily make a trade-off between the gain and the noise figure. As to say for a given
noise figure, the point at which a constant gain circle (with maximum possible gain) is
406 Chapter 9. Amplifier Design Using S-parameters
īL $ĺ0LQQRLVHILJXUH)PLQ
&RQVWDQW
JDLQ %ĺ0D[SRZHUJDLQ
FRQWRXUV
G%
% &RQVWDQW
1)
&
FRQWRXUV
$
G%
G%
G%
G% G%
G%
G%
īU
G%
G%
G%
ī$
ī%
ī&
Figure 9.15: Constant NF and the normalized available power gain contours on
the input reflection coefficient plane at 6 GHz.
tangent to it gives the best compromise, otherwise for a given power gain the point at
which a constant noise figure circle (with minimum possible noise figure) is tangent
to it gives the best compromise. For example, if we choose to have a NF of 1 dB, we
would then choose point C where the 0 dB power gain circle is tangent to it and the
corresponding ΓS is 0.45∠47◦ .
S21 S12 Γs ∗
ΓL = S22 + (9.124)
1 − S11 Γs
Example 9.5 The optimum noise bias point for a given bipolar transistor is VCE =
10 V and IC = 5 mA. The optimum noise input reflection coefficient at 200 MHz is
Γs = 0.7∠140. S-parameters at 200 MHz (in a 50 Ω measurement system) are given
as S11 = 0.4∠168◦ , S12 = 0.04∠60◦ , S21 = 5.2∠63◦ , S22 = 0.35∠ − 39◦ . Design
an LNA at 200 MHz with the source and load impedances of 75 Ω and 100 Ω. Then,
determine what the transducer power gain is expected of this design.
Solution:
We first evaluate the Rollet stability factor as
Considering k > 1 and |∆| < 1 the transistor is unconditionally stable. As such, the
amplifier is stable. We then design the input matching network for a 75 Ω source
impedance. We depict the optimum noise source impedance on the Smith chart
at point C, Γs = 0.7∠140 corresponding to ZnC = 0.2 + j0.35. Using Figure 9.16,
starting from point A (ZnA = 1.5 + j0), we turn clockwise on the constant conduc-
tance contour (G = 0.667) to intersect the constant resistance contour (R = 0.2) at
point B. So for the parallel capacitor, we would have
Γi
Γs, opt
C
A
Zsource=75Ω
Γr
YnA=0.667+j0
B
YnB=0.667+j1.7
YnC=1.23-j2.153 ΓC=ΓS=0.7 +140º
ZnA=1.5+j0
ZnB=0.2-j0.51
ZnC=0.2+j0.35
Γi
B
Γmatching
A
Zload=100Ω
Γr
YnA=0.5+j0
YnB=0.5-j0.48
ZnA=2+j0 ΓB=ΓL=0.43 +61º
ZnB=1.04+j1
Starting from point B (ZnB = 0.2 − j0.51), we turn clockwise on the constant
resistance contour (R = 0.2) to arrive at point C. So for the series inductor, we
would have
|Im{ZC } − Im{ZB }| 50 × 0.86
L= = = 34nH (9.127)
ω 2π 200 × 106
1 1
L= = = 83 nH (9.129)
ω |Im{YC } − Im{YB }| 2π 200 × 106 0.48 × 0.02
The 330 pF capacitors at the input and the output are for DC decoupling purpose
and their AC impedances are negligible.
9.8 Design of Two-Stage Amplifiers 409
9
S) ȍ
N Q+
Or
Now that we have learned how to design a single-stage amplifier, we move on to design
a two-stage low noise amplifier and understand the design procedure.
Example 9.6 If the amplifier whose specifications are given in Figure 9.15 has a
value of |S21 | = 8 and it is cascaded by another amplifier whose NF is F2 = 5 dB,
find the overall NF once the amplifier’s source reflection coefficient is either at
point A or at point B.
Solution:
We record NF and the power gain from Figure 9.15 as F = 0.8 dB and G =
−1 dB at point A and F = 4 dB and G = 3 dB at point B. The denormalized power
gains at points A and B become
GA = −1 + 10 log 8 = 8 dB (9.133a)
GB = 3 + 10 log 8 = 12 dB (9.133b)
This design procedure for a specific power gain and the minimum possible NF consists
of three steps: (1) The contours corresponding to the desired APG should be drawn;
(2) The NF contour which is tangent to the required power gain contour should be
drawn next; and (3) The optimum reflection coefficient lies on the specified loci where
the two circles are tangent.
Example 9.7 The optimum bias points along with S-parameters and noise pa-
rameters for◦
a transistor are ◦provided at 4 GHz:
◦
VCE = 10V , IC =◦ 4 mA, S11 =
0.552∠169 , S12 =◦0.049∠23 , S21 = 1.681∠26 , S22 = 0.839∠−67 , Fmin = 2.5 dB,
Γopt = 0.475∠166 , rn = 3.5 Ω. Design the amplifier such that the overall noise
figure of this stage followed by another stage with NF = 7 dB is minimized.
Solution:
We first evaluate the stability of the transistor.
As a result, the amplifier is unconditionally stable. The constant APG and the
constant NF contours are plotted in Figure 9.19 in the plane of the source reflection
9.8 Design of Two-Stage Amplifiers 411
coefficient. Now, we choose a set of points with the specified noise figure and
the maximum possible APG as follows, and we compute the overall noise figure
consequently (from Equation 9.132).
2.5dB
A1
A2
Γr
A3
A4
14.7dB
Constant
gain
contours ΓA2=ΓS=0.5 -175º
Figure 9.19: Constant power gain and constant NF contours for the given
transistor on the Smith chart.
Γs = 0.524∠+186◦ (9.142)
For the output matching, the load reflection coefficient can be calculated as
S21 S12 Γs ∗
ΓL = S22 + = 0.871∠+70◦ (9.143)
1 − S11 Γs
412 Chapter 9. Amplifier Design Using S-parameters
It is noteworthy that all the procedure which has been described in terms of S-
parameters formulation can be repeated in terms of other circuit parameters such as
Y-parameters. The following example illustrates this approach.
Example 9.8 Admittance parameters for a field effect transistor at 1 GHz are
given as Y11 = 8.79 mjf,Y12 = −2.5 mjf,Y21 = 1 mf − 2.5 mjf,Y22 = 0.33 mf +
3.77 mjf.
(a) Determine the circuit model values as depicted in Figure 9.20.
(b) Does this transistor lead to a stable amplifier design? Determine the value of
the required parallel resistor at the output to provide unconditionally stability if a
parallel 1 kΩ is added at the input.
(c) Assuming the output is short circuited, for matching the input to 50 Ω, a π
section is used, as shown in Figure 9.21. Determine the values of the capacitors
and the inductor with an input quality factor of 10.
(d) Calculate the matching bandwidth.
&JG
50Ω C2
+
Vs C1 L 1kΩ Cgs Cgd
Solution:
(a) The equivalent model of the transistor can be shown as in Figure 9.20. The
values can be found using the provided Y-parameters as
Y12 = − jωCgd ⇒ Cgd = 398 fF (9.144)
Y11 = jω Cgs +Cgd ⇒ Cgs = 1 pF (9.145)
Y21 = gm − jωCgd ⇒ gm = 1 mf (9.146)
Y22 = gds + jω Cgd +Cds ⇒ gds = 330 µf,Cds = 202 fF (9.147)
(b) The Rollet’s stability factor in terms of Y-parameters can be expressed as [5]
2g11 g22
k= (9.148)
|Y12Y21 | + Re{Y12Y21 }
9.8 Design of Two-Stage Amplifiers 413
Given the fact that g11 = 0 then k = 0 which suggests instability. Adding a 1 kΩ
resistor at the input, the Rollet’s stability factor can be rewritten as [5]
This indicates that the circuit is unconditionally stable in this case, and no parallel
resistor is needed at the output.
(c) The value of the components can be found as
n2
Q= Ceq ω = 10 (9.150)
GS
n2
Let, GS = 1k, then Ceq = 1.59 pF, where
r
C1 +C2 1000
n= = = 4.47 (9.151)
C2 50
and
C1C2
Ceq = (9.152)
C1 +C2
1
= Ceq ω + b11 = 10 m + 8.79 m = 18.79 mf (9.154)
Lω
L = 8.47 nH (9.155)
(d) For estimating the matching network bandwidth, we calculate the total equiva-
lent resistance in parallel with the inductor
Example 9.9 Assume that the bandwidth of a receiver is equal to 30 kHz, and the
maximum allowable noise figure at the input is 7 dB. If the required SNRmin = 6 dB,
compute the sensitivity of the receiver.
Solution:
We have
9.9 Conclusion
In this chapter, we discussed the design of RF/microwave amplifiers. We investigated
the stability condition at the input and the output in terms of S-parameters. We learned
about the radii and the centers of the stability circles at the source and at the load
reflection coefficient planes. After resolving the stability problem, we presented three
distinct definitions for the power gain of a two-port amplifier which were the operating
power gain, the APG, and the transducer power gain. Then, we introduced the constant
operating power gain contours on the load plane and the constant APG contours on
the source plane over the Smith chart. Subsequently, we provided the definition of the
noise figure and introduced constant NF contours on the source plane. Finally, a 3-step
design procedure was provided by means of which the trade-off between noise figure
and the power gain can be achieved, leading to the low-noise design with proper power
gain.
9.11 Problems
Problem 9.1 Prove that in the case where the source and the load impedances are
equal to the characteristic impedance, the transducer power gain (GT ) can be written
as GT = |S21 |2 . Then calculate the operating power gain (GP ) and the APG (GA ) in
terms of the S-parameters of the transistor.
Problem 9.2 Consider the circuit shown in Figure 9.22. Compute GT , GA , and GP ,
with the following parameters:
Γs = 0.49∠ − 150◦ , ΓL = 0.56∠90◦
S11 = 0.54∠165◦ , S12 = 0.09∠20◦ , S21 = 2∠30◦ , S22 = 0.5∠ − 80◦
Output
50Ω Input
matching 50Ω
+ matching
network
Vs network
ΓS ΓL
Figure 9.22: The amplifier circuit for determining various power gains.
Problem 9.3 The S-parameters for three transistors are given. Comment on their
stability by drawing the stability circles at the source and the load planes.
◦ 0 ◦ 00 ◦
S11 = 0.674∠ − 152 S11 = 0.385∠ − 55 S11 = 0.7∠ − 50
◦ 0 ◦ 00 ◦
S12 = 0.075∠6.2 S12 = 0.045∠90 S12 = 0.27∠75
◦ 0 ◦ 00 ◦
S21 = 1.74∠36.4 S21 = 2.7∠78 S21 = 5∠120
◦ 0 ◦ 00 ◦
S22 = 0.6∠ − 92.6 S22 = 0.89∠ − 26.5 S22 = 0.6∠80
Problem 9.4 Show that as S12 approaches zero, the centers and the radii of the input
and output stability circles can be estimated as CS ≈ S111 , rS ≈ 0,CL ≈ S122 , rL ≈ 0.
Given |S11 | < 1 and |S22 | < 1, what would you deduce from this?
Problem 9.5 Two different amplifiers with the specified S-parameters are cascaded
as shown in Figure 9.23. Compute the overall S-parameters of these cascaded amplifiers
in terms of their corresponding S-parameters (Hint: use the concept of loaded two-port
S-parameters).
D D
E 6$ 6$ 6% 6% E
3RUW 3RUW
6$ 6$ 6% 6%
Problem 9.6 Consider the circuits illustrated in Figure 9.24. Determine how resistive
loading affects the overall S-parameters and the stability of the two-port network. The
transistor’s S-parameters are:
S11 = 0.69∠ − 78◦ , S12 = 0.033∠41.4◦ , S21 = 5.67∠123◦ , S22 = 0.84∠ − 25◦
9.11 Problems 417
Hint: (1) Use the results of problem 9.5, (2) For the series resistance two-port S-
parameters, one can easily show that S21 = 1 − S11 , and for the parallel resistance
two-port S21 = 1 + S11 .
29Ω
9Ω
500Ω
71.5Ω
(a) (b) (c) (d)
12V
RC 1nF
70nH
RB
1nF 30pF
β=100 70pF
50Ω
50Ω
+ 40nH
Vs
ΓS ΓL
Problem 9.12 The S-parameters and the noise parameters of a transistor are given at
1 GHz as:
S11 = 0.6∠170◦ , S12 = 0.05∠16◦ , S21 = 2∠30◦ , S22 = 0.5∠ − 95◦
Fmin = 2.5 dB, Γopt = 0.5∠145◦ , Rn = 5 Ω
Verify the stability and determine the maximum GA . Then, plot the constant power
gain contour having a gain 3 dB lower than GA,max . Furthermore, plot the constant NF
contours for NF = 3 dB and NF = 4 dB. Finally, derive the NF of the amplifier at the
maximum power gain point (in the source impedance plane).
Problem 9.13 Consider Figure 9.26. We wish to design a 2 GHz amplifier having
NF = 2 dB with maximum possible GT . First determine the required ΓS and corre-
sponding GA . Then compute the required Γout . What would be the value of GT then?
The S-parameters are given at 2 GHz as:
S11 = 0.646∠172◦ , S12 = 0.051∠13.5◦ , S21 = 3.042∠47.9◦ , S22 = 0.642∠ − 64◦
Constant
Γi
NF
contours
4dB
3.5dB
3dB
2.5dB
2dB
1.7dB
Γr
16dB
15dB
14dB
12dB
10dB
Constant
available gain
contours
Problem 9.15 For the circuit depicted in Figure 9.28, calculate the input and the
output power in dBm as well as the output voltage in dBVolts.
ZIN=50Ω ZOUT=50Ω
50Ω
Es,rms= + GP1=13dB GP2=13dB GP3=10dB GP4=9dB 50Ω
79.5mV
Problem 9.16 For the MOS amplifier operating at 2 GHz whose S-parameters are
S11 = 0.3∠160◦ , S12 = 0.03∠62◦ , S21 = 6.1∠65◦ , S22 = 0.4∠ − 38◦ , and depicted in
Figure 9.29
1. Neglecting S12 determine the input and the output matching loads and calculate
Gp . Assuming ideal inductors and capacitors, determine their required values.
2. If L3 is replaced by a short-circuited 50 Ω stub, calculate the electrical length of
the line as a fraction of the wavelength.
VG VDD
RFC L3
C4
50Ω C∞ L2
+ 50Ω
vS C1
Figure 9.29: The transistor MOS amplifier for input and output matching.
420 Chapter 9. Amplifier Design Using S-parameters
G1 G2
BGU7007 NF2=5dB
G2=14dB
choose a point on the contours depicted in Figure 9.31 to have the minimum
noise figure in this case. Determine the required load reflection coefficient to
match the output, and then design the matching network in a 50 Ω system at
both the input and the output using air-filled T-lines (εr = 1) and open/short
stub.
Γi
Constant
NF
contours
N1
N2
N3
N4
N5
Γr
G1
G2
G3
G4
G5
G6
G7
G8
G9
Constant
gain
contours
Gmax=14.7dB; G1=14.5dB; G2=13.7dB; G3=12.7dB; G4=11.7dB;
G5=10.7dB; G6=9.7dB; G7=8.7dB; G8=7.7dB; G9=6.7dB;
Figure 9.31: The constant noise and the constant available gain circles at the
source plane of the transistor at 4 GHz.
Problem 9.21 Calculate the maximum GT , if stable, for the Motorola silicon bipolar
transistor with part number MRF962 at 700 MHz at different biases of VCE = 5V, IC =
10mA, 25mA, 50mA where the S-parameters data are as follows:
Problem 9.22 Evaluate stability at 1.5GHz and for VCE = 5V, IC = 25mA, and design
the matching network for maximum transducer gain at this frequency. Use Table. 9.1.
422 Chapter 9. Amplifier Design Using S-parameters
Problem 9.23 In the given amplifier depicted in Figure 9.32, match the input and
the output at the frequency of 318.3 MHz. First determine the input and the output
admittances of the transistor at the operating frequency, and then find the required
reactive components to match the input to 50 Ω and the output to 20 Ω.
L1 Lg rg rd Ld L2
50Ω
C1 Cgs gmv rds Cds C2 20Ω
rds=400Ω
Cgs=15pF
Cds=5PF
rg=12Ω gm=50mS
Ld=2.0nH
Lg=2.5nH
rd=5Ω
Figure 9.32: The equivalent circuit of a FET transistor and the associated
matching circuits.
Problem 9.24 The Y-parameters of the cascode MOSFET stage at 159 MHz are
depicted in Figure 9.33. Determine the reactive matching components for the maximum
transducer power gain and compute the mentioned power gain.
L3
L1
RS
L2 RL
+ C1
Vs
yi=2.6+j12 ms
yr=0 RS=RL=75Ω
yf=-56+j14 ms
yo=1.1+j2.1ms
Rn h i
F = Fm + (Bs − Bo )2 + (GS − Go )2 (9.161)
GS
9.11 Problems 423
Bs(ms)
-6 2.5+j5 0
F=3dB Y T= mS
-5
52-j21 1+j2.5
-4
-3
-2
-1 Gs(ms)
1 2 3 4 5 6 7
L1 L2
YT
RS
C1 C2 RL
Figure 9.34: The low-noise amplifier with the corresponding input and output
matching circuits.
Problem 9.26 In a low-noise amplifier, a noise figure of F = 3 (or 5 dB) has been
measured for three consecutive source admittances, YS1 , YS2 , and YS3 as depicted in
Figure 9.35. Draw the corresponding constant noise figure circle and from there
determine the corresponding optimum noise admittance. Use Equation 9.161.
Bs(ms)
-60
YS1=37.3-j20 mS
-50
YS2=2.7-j20 mS
-40
YS3=20-j2.7 mS
-30
-20
-10 GS
10 20 30 40 50 60 70
Figure 9.35: The measured source admittances for a noise figure F = 3 (source
admittance plane).
ID
VGS
RF power amplifiers (PAs) consume the highest amount of power among all the
transmitter blocks. While advancements in technology have resulted in aggregating
all transmitter blocks into one single integrated circuit, the PA block is still integrated
separately in many applications. In a transmitter chain, the data signal modulates
properly the carrier signal and is then upconverted from the IF to the RF frequency.
Afterward, a PA provides the necessary RF power level to transmit the signal according
to the standard of interest, and the signal is radiated into the air by the antenna. By
virtue of the power amplification, the PA mostly operates in a nonlinear or large-
signal regime which mandates careful considerations regarding both its design and
its simulation. One of the most important trade-offs in a PA is that of efficiency and
linearity, as shown in Figure 10.1. In customary PAs, more linearity is expected to
result in less efficiency and vice versa, while better linearity and good efficiency are
both required to obtain a high data rate and low power consumption, respectively.
PA nonlinearity stems from the large-signal behavior of the active devices. An
important issue that must be taken into consideration is the presence of the signal har-
monics as well as IM products which have a potentially adverse effect on the adjacent
channels. For this reason, the standards stipulate a measure called adjacent channel
power ratio (ACPR) to regulate this issue. Furthermore, efficiency considerations im-
pose lower limits on the supply voltage and active device architecture. In other words,
the voltage source must be capable of providing sufficiently high currents and the active
device should have a high voltage swing without entering the breakdown region. For
driving purpose, a predriver stage is often used before the PA to provide the required
signal level at the PA input. Noting the fact that the antenna impedance is in the order
of 50 Ω and the breakdown voltage limit in active devices, an impedance matching
circuit is utilized to match the required load impedance to the antenna impedance
level. Considering the limited supply voltage for higher powers, we would need higher
current swings which means lower impedance levels at the PA output. Impedance
matching circuits generally introduce a finite loss due to use of low-Q inductors or
capacitors, and consequently causing a poorer efficiency.
426 Chapter 10. Power Amplifier
(IILFLHQF\
HQKDQFHPHQW /LQHDUDQG
/LQHDULW\
HIILFLHQW3$
/LQHDUL]DWLRQ
&RQYHQWLRQDO
3$V
(IILFLHQF\
10.1 PA Specification
In this section, the most vital design specifications of PAs are introduced and studied.
These specifications consist of the efficiency, the output power, the in-band noise, the
gain, the linearity (AM to AM and AM to PM distortion, ACPR, and error vector
magnitude (EVM)), and the stability. In general, power amplification can be discussed
in two categories dependent on the linearity of the operating region. Assume the PA as
a block with a single-tone input as
where G is the PA gain. Unlike the linear amplification, the output signal in a nonlinear
amplifier will take the form of
10.1.1 PA Efficiency
Efficiency is the most critical parameter in PA design. A PA with 50% efficiency
delivering 1W power (or 30 dBm) to a 50 Ω load, for example, also dissipates 1 W
10.1 PA Specification 427
power in the circuit, and thus having a total power consumption of 2 Watts. The
dissipated power generates heat and necessitates specific measures in the circuit
implementation, also shortening the battery life. In the context of PA design, power-
added-efficiency (PAE) and efficiency are defined as
Pout,RF − Pin,RF
PAE = (10.4)
Ptotal,DC
Pout,RF ∼ Pout,RF
η= = (10.5)
Ptotal,DC + Pin,RF Ptotal,DC
In Equation 10.4, the numerator accounts for the difference between high-frequency
output power and the input power, and the denominator represents the total DC power
dissipation. Equation 10.4 gives the power transferred to the load minus the input power
divided by the DC power, and is called PAE for that matter. As studied in previous
chapters, the output power exhibits a compressive behavior as a function of the input
power as shown in Figure10.2. In other words, owing to nonlinear amplification,
the output signal no longer increases in proportion to the input signal when the input
amplitude exceeds a certain value. The rate of change of the numerator in Equation 10.4
becomes smaller, leading to a compression in PAE as shown in Figure 10.3.
As an example for the importance of PAE in cell phones, the higher the PAE of a
cell phone’s PA, the longer would be the battery lifetime and the call durations.
Pout
(dBm)
Compression curve
30
25
20
15
Pin
-10 -5 0 5 10 (dBm)
3RZHUDGGHG
HIILFLHQF\
3LQ
G%P
PSPEC+L
L PSPEC
PA
The band-pass filter (BPF) is used to suppress the spurs and the undesired har-
monics generated by the PA, also providing a proper load impedance for the amplifier.
Nevertheless, the ultimate efficiency is degraded due to the insertion loss of the filter.
To alleviate this issue, the PA must be designed for higher output power and better
efficiency. For Figure 10.4, PAE can be written as
Pout ∗ L − Pin
PAE = (10.6)
Ptotal,DC
Here, L is the insertion loss of the output filter. Note that the numerical value of
L should be used in the above equation instead of its dB value. Each communication
standard allows a specified amount of power to be transmitted. The effective radiated
power (ERP) can be defined as the product of the power supplied to the antenna by the
antenna gain relative to a half-wave dipole in a given direction. Another definition also
exists as equivalent isotropically radiated power (EIRP) which is the product of the
power supplied to the antenna by the antenna gain in a specific direction relative to an
isotropic antenna. In general
where the 2.2 dB term is the gain of half-wave dipole antenna with respect to an
isotropic antenna.
Table 10.1 shows various specifications of output power for a number of commu-
nication standards.
Another parameter of the PA is characterized with its probability density function
(PDF). To be adapted to the output power requirements (e.g., in a CDMA system), the
10.1 PA Specification 429
3')
6XEXUEDQ
S3
8UEDQ
2XWSXW
SRZHU
3G%P
Figure 10.5: PDF of the transmitted power of the PA for a mobile set in an
urban or suburban area.
mobile unit mostly transmits a power inferior to the maximum value. For this reason,
in urban areas with more base stations, the chances of successful transmissions are
higher even at low power levels, while power should be necessarily high in suburban
areas. The PDF depicts the probability distribution of the power transmitted from a
CDMA transmitter unit, a sample of which is shown in Figure 10.5 for an amplifier in
an urban area or in a suburban area.
%6WUDQVPLW
5HFHLYH
VLJQDO I7; I5; I7; I5; I7; I5;
VLJQDO
I7; I5;
/1$
'XSOH[HU
%DVH
VWDWLRQ 7UDQVPLW
VLJQDO
3$
Figure 10.6: The effect of the TX-band noise at the receiver in a full-duplex
system.
duplexer provides about 30 to 40 dB isolation between the receive and the transmit
paths the reason for which the transmitted signal could degrade the receiver sensitivity
by leaking to it. In fact, the PA amplifies the noise in the receiver band as well as
the desired signal in the transmitter band. The PA noise itself will be added to the
receiver, as such degrading its performance. Figure 10.6 demonstrates this process.
In Figure 10.6, the base station transmits the signal with a large output level and the
attenuated version of the original signal is received at the receiving antenna. The PA
amplifies the signal in the transmitter band, but it leaks to the receive path through the
duplexer (due to finite isolation between the two paths). In addition, the PA noise in
the receiver band is only slightly attenuated by the duplexer and is added to the receive
path. Consequently, detecting the received signal will be tougher, i.e, the receiver
sensitivity will be degraded.
10.1.4 PA Gain
In practice, the minimum required PA gain is affected by four parameters:
• Maximum output power
• Loss after the PA
• Maximum driver output power
• Loss after the driver
Now, consider the system in Figure 10.7 in which a driver precedes a PA and there are
interstage matching networks with a certain loss.
The overall minimum gain of the power amplifier in Figure 10.7 then can be
obtained as
Gainmin (dB) = PSPEC,out (dBm) + L(dB) + Ld(dB) − Pmax Driver (dBm) (10.8)
10.1 PA Specification 431
Figure 10.7: A PA with the PA driver and the interstage matching networks.
26
24 Pout,Max L
Pout(dBm)
22 PSPEC
20
18
16
Pin,Max
14
-10 -8 -6 -4 -2 0 2 4
Pin(dBm)
Figure 10.8: Typical compression of the output power versus the input power
of a PA.
26
24
Gain (dB)
22
20 Gain compression
18 at peak power
16
Pin,Max
14
-10 -8 -6 -4 -2 0 2 4
Pin(dBm)
3RXW 3RXW
G%P G%P
6ORSH
RI
2,3 6ORSH
RI
&,
G% 2XWSXW
FRPSUHVVLRQ
3LQG%P I
WK WK UG I I UG WK WK
3LQG% ,,3
D E
Figure 10.10: (a) P1dB and IP3 points in a nonlinear system, (b) Output spectrum
in a nonlinear system with a two-tone input.
10.1 PA Specification 433
Vout
+
vS R GmvS R CD(V1)
-
vS=VScos(ωt)
where
GmVs R
V1 = q (10.14)
1 + (RωCD (V1 ))2
By substituting the value of CD (V1 ) from Equation 10.12, one can rewrite Equa-
tion 10.14 in the following form
2
3 3
RωC3 V16 + (Rω)2 C3C1V14 + 1 + (RωC1 )2 V12 − (GmVS R)2 = 0 (10.15)
4 2
By resolving the above equation, one can obtain the value of V12 , and therefore obtain
the phase value as
φ (t) = −tan−1 (RωCD (V1 )) (10.16)
According to Equation 10.16, the output phase is a function of the input voltage
amplitude. Figure 10.12 shows a sample of phase shift in the output of a system as the
input signal amplitude increases. As the input signal power is raised, the output phase
begins to change significantly after passing a threshold power.
It should be noted that phase modulation results in spectral regrowth. To gain
a better understanding of spectral regrowth, consider the output signal of a phase-
modulated single-tone sinusoid as
Vout(t) = A cos(ωct + k sin(ωbt))
= A cos(ωct) cos(k sin(ωbt)) − A sin(ωct) sin(k sin(ωbt)) (10.17)
434 Chapter 10. Power Amplifier
AM to PM curve
-10
Φout(degree)
-12
-14
-16
-18
-20
-22
-10 -8 -6 -4 -2 0 4
Pin(dBm)
Figure 10.13(a) shows the one-sided frequency spectrum of Equation 10.18 for k 1,
only the first harmonic of the baseband will appear about the carrier. In the case
where k ≈ 1 or k > 1 (Figure 10.13(b)), the harmonics of ωb will also appear in both
sidebands and we can observe a certain spectral regrowth where in Equation 10.17,
the carrier multipliers cos(k sin(ωbt)) and sin(k sin(ωbt)) can be expressed in terms of
Bessel functions of even and odd order, respectively
As such, the even and the odd harmonics of the baseband signal will also modulate the
in-phase and the quadrature components of the carrier and the signal spectrum will
grow. In the following sections, some effects of nonlinearity will be briefly discussed:
Generation of undesired harmonic components: High gain and high output
voltage swing limitations lead the PA to perform in a large-signal regime. This results
Vout Vout
k<<1 k<1 or k≈1
ω ω
ωc-3ωb
ωc-2ωb
ωc
ωc-ωb
ωc+ωb
ωc+2ωb
ωc+3ωb
ωc
ωc-ωb
ωc+ωb
(a) (b)
where Idn is the current component of the nth harmonic of the main frequency. Further-
more, if two or more blockers are present at the input of the nonlinear system, the IM
causes more spurious components to be generated (Figure 10.14(b)). This is why using
filters in the output of PAs is common to attenuate the undesired harmonic components.
Spectral regrowth: The nonlinear performance of the system results in the gener-
ation of undesired components in the spectrum, having adverse effects on the adjacent
channel. Both AM to AM and AM to PM conversions cause the spectrum to regrow
at the output of the amplifier. Adjacent channel power ratio (ACPR) represents the
amount of spectral regrowth in a PA, thus being a criterion of its linearity. Imagine
a nonlinear tuned amplifier as depicted in Figure 10.15 where its nonlinear dynamic
transconductance is described as follows
i = αv + β v3 + γv5 (10.22)
Here
3
3 Ȧ
Ȧ 9LQ 9RXW Ȧ
Ȧ
Ȧ
Ȧ Ȧ
'&
D
3
3
Ȧ
Ȧ
ȦȦ
ȦȦ
ȦȦ
ȦȦ
ȦȦ
ȦȦ
Ȧ Ȧ
Ȧ
Ȧ
9LQ 9RXW
Ȧ
Ȧ
Ȧ Ȧ
'&
E
Vout
+
vS2 +
-
+ v i=f(v) RL CL LL
vS1 -
-
Then
i =α (vS1 + vS2 )
+ γ v5S1 + 5v4S1 vS2 + 10v3S1 v2S2 + 10v2S1 v3S2 + 5vS1 v4S2 + v5S1 (10.24)
Putting the sinusoidal voltage terms in Equation 10.24, we can compute the output
fundamental voltages as well as the third and the fifth IM terms which happen to be
within the output bandwidth. The third and fifth IM currents will have the following
forms
2V
3βVS1 S2
I2ω1 −ω2 = cos ((2ω1 − ω2 )t) (10.25a)
4
2
3βVS1VS2
I2ω2 −ω1 = cos ((2ω2 − ω1 )t) (10.25b)
4
3 V2
5γVS1 S2
I3ω1 −2ω2 = cos ((3ω1 − 2ω2 )t) (10.26a)
8
2 V3
5γVS1 S2
I3ω2 −2ω1 = cos ((3ω2 − 2ω1 )t) (10.26b)
8
The output voltage including the IM terms (considering flat impedance, RL , for all of
the terms) can be described by the following expression.
As it is seen here when there is at least two carrier frequency components at the
input, the output spectrum will be widened by a factor of three due to the third-order
IM and it will be widened by a factor of five due to the fifth-order IM. This results
in a spectral regrowth at the sidebands of the carriers as seen in Figure 10.16. Also,
Figure 10.17 depicts spectral regrowth and the resulting distortion in the time domain.
Note that while the fundamental terms grow with a slope of 10 dB/decade, the
,QSXWVLJQDO
6LJQDO
2XWSXWVLJQDO
%:
3RZHUVSHFWUDOGHQVLW\
G%P+]
)UHTXHQF\+]
Aout1
Ain1
t t
Vin Vout
f Linear PA f
123 123
Aout1
Ain1
t t
Vin Vout
f Nonlinear PA f
123 123
Figure 10.17: Spectral regrowth in a nonlinear PA, single tone shown in the
time domain.
438 Chapter 10. Power Amplifier
third-order IM terms grow with a slope of 30 dB/decade, and the fifth-order IM terms
grow with a slope of 50 dB/decade. This point indicates that the input of the power
amplifiers should not be more than a specified level in order to maintain the required
signal to IM ratio or the required ACPR.
For instance, in CDMA (IS-95), ACPR is defined as the ratio of the available
power at 1250 kHz offset frequency (with respect to the carrier) in a 30 kHz bandwidth
to the available power in the main channel (at the center frequency) within the same
bandwidth, as shown in Figure 10.18. Therefore, ACPRLo and ACPRHi can be defined
as
PLo
ACPRLo = (10.28a)
Po
PHi
ACPRHi = (10.28b)
Po
3R
3/R 3+L
)UHTXHQF\0+]
Care must be taken with regard to the above method as it is completely an approxi-
mative one. So its result might underestimate the spurious generated in the adjacent
channels (sidebands), because these unwanted components are mainly generated due
to a dynamic IM process.
Decreasing the noise power ratio: The nonlinear performance of PAs increases
the in-band noise power especially at the carrier frequency as well as generating
undesired components at the adjacent channels. This causes a lower SNR and thus
lower sensitivity in the receiver. In other words, the transmitted data will be noisy and
hard to be detected by the target receiver.
One of the ways to determine in-band distortion is measuring the noise power ratio
(NPR). For this purpose, the input signal is initially passed through a very sharp notch
filter to ensure that its in-band spectrum has no components at the carrier frequency.
Then, the resulted signal spectrum is applied to the nonlinear system as an input. The
nonlinear performance of the system will lead to spectral regrowth and increasing the
intermodulation noise floor level at the carrier frequency. Ultimately, the ratio of the
obtained IM noise power at the carrier frequency to the output signal power in the band
is calculated as NPR, as shown in Figure 10.19.
Increase in Error Vector Magnitude: Error vector magnitude (EVM) is defined
as the normalized distance between the desired and actual signal vectors as depicted in
Figure 10.20. Due to the nonlinear performance of the system, the yielded vector is
not expected to coincide with the desired signal vector in practice.
EVM is reported in both rms and peak values. Typical values range from 7% to
12% in the former and from 22% to 33% in the latter.
Increases in EVM can be interpreted as distortion in the data constellation at
the output signal. Undesired changes in phase and amplitude bring about errors in
detecting the modulated data, causing problems in the data transmission process. This
issue becomes more significant when the modulation level is higher, such as in QAM.
Figure 10.21 depicts a sample of data constellation distortion at the output for 16-QAM
modulation.
3RZHUVSHFWUDO
,QSXWVLJQDO 2XWSXWVLJQDO
GHQVLW\
GHQVLW\
135
I& I&
)UHTXHQF\ )UHTXHQF\
4 4
, ,
7\SLFDOLQSXW 7\SLFDORXWSXW
FRQVWHOODWLRQ FRQVWHOODWLRQ
Figure 10.21: A comparison of the input and the output constellation for a
16-QAM modulation for a nonlinear power amplifier in between.
10.2 PA Topologies
The PA topologies have evolved through the recent decades. Differences in linearity
and efficiency are the main reasons that impose various topologies. Another important
parameter that makes a significant difference in PA performance is the conduction
10.2 PA Topologies 441
angle which varies depending on the operation point of the active device. In addition,
as the efficiency is of great importance in the design of a PA, a method must be sought
to enable minimum power dissipation by the active devices. Harmonic termination can
be helpful in some PA classes. In the next sections, the details of the operation classes
of PA will be studied as well as the relations regarding the efficiency of each class.
VG VDD
id Conduction
RFC RFC angle 360º
C∞ Vout
Vd
ZS C∞
Vin IDC
RL
VS ωt
π
2π
3π
4π
Input matching
network
id
Id,max
Ideal
transfer IDC
ID function
t
Quiescent vd
point
Vd,max
VDD
Cut-off
region VTH VGS
t
Output PLoss
waveform
PLoss in the device
Input
waveform
t
(a) (b)
Figure 10.23: (a) Quiescent point determination in class A power amplifiers and
(b) Drain voltage, drain current, and device power loss of the circuit depicted
in Figure 10.22.
VDD
id Conduction
RFC angle 180º
C∞ Vout
Vd
ZS
Vin
RL
VS ωt
π
2π
3π
4π
Input matching
network
Figure 10.24: A class B power amplifier with the corresponding output current
waveform.
444 Chapter 10. Power Amplifier
id
ID2
Id,max
t
vd
Quiescent
point VDD
VTH VGS
t
PLoss
PLoss in the device
t
(a) (b)
Figure 10.25: (a) The quiescent operating point and the corresponding wave-
forms for the class B PA and (b) The output current, the output voltage, and the
power loss waveforms for a single-device class B power amplifier.
Finally, noting that the drain voltage in Figure 10.24 can maximally swing from zero to
the supply voltage, the maximum efficiency is obtained by the ratio of the fundamental
(main harmonic) power to the consumed DC power by the source:
1 VDD Imax
Pac,max 2 2 2 π
ηmax = = VDD Imax
= ≈ 78.5% (10.42)
PDC 2 π
4
Note that in this case, for maximum efficiency, the load resistance should be chosen as
VDD
2 VDD
RL = Imax
= (10.43)
2
Imax
VDD
RL = (10.45)
2Imax
Therefore, the load impedance in the class B push–pull case would be half the load
impedance in a single-ended class B power amplifier (with the same device and the
same bias).
id1
VDD
M1 id2 ωt
ZS Vout
Vin
VS M2
RL
iout ωt
ωt
π
2π
3π
4π
Figure 10.26: A class B push–pull power amplifier with the output current
shown as a summation of the positive and the negative half-cycle currents.
446 Chapter 10. Power Amplifier
while its linearity is better than class B and worse than class A. Figure 10.27 shows the
quiescent point determination curve for class AB amplifiers in general.
Thus, the average current (DC current) will be obtained by integrating the current
waveform over one period:
Z T Z
1 1 θ2
I0 = Iav = id (t) dt = Ip sin (ωt) − ID d (ωt) (10.47)
T 0 2π θ1
ID
Quiescent
point
id Conduction
angle >180º
2π
3π
4π
Figure 10.27: Class AB PA’s quiescent point and the corresponding current
waveform.
10.2 PA Topologies 447
VG<0 VDD
id Conduction
RFC RFC angle <180º
C∞ Vout
Vd
ZS C∞
Vin 2θ
RL
VS ωt
θ1
θ2
π
2π
3π
4π
Input matching
network
Figure 10.28: A class C power amplifier and its corresponding output current
waveform.
id
ID2
t
vout
Quiescent
point
VTH VGS
t
PLoss
PLoss in the device
t
(a) (b)
Figure 10.29: The quiescent point of a class C amplifier and the corresponding
output current, output voltage, and the power loss waveforms.
where 2θ equals the conduction angle (α) of the PA. The total DC power consumption
will be given by
IP
PDC = VDD I0 = VDD (sin (θ ) − θ cos (θ )) (10.49)
π
To obtain the AC power delivered to the load, the fundamental current is calculated as
Z π +θ
1 2 Ip
I1 = Ip sin (ωt) − ID sin (ωt) d (ωt) = (2θ − sin (2θ ))
π π −θ
2
2π
(10.50)
Finally, assuming the peak output AC voltage as VDD , the efficiency of the class C
amplifier is yielded as
1 I
PO VDD p (2θ − sin (2θ )) 2θ − sin (2θ )
ηmax = = 2 I 2π = (10.51)
PDC VDD πP (sin (θ ) − θ cos (θ )) 4 (sin (θ ) − θ cos (θ ))
where 0 < θ < π. It should be noted that for maximum AC power and efficiency, we
should choose the following value for the load impedance.
VDD VDD VDD 2π
RL = = IP
= (10.52)
I1 2π (2θ − sin (2θ )) IP (2θ − sin (2θ ))
It is helpful to note that Equation 10.51 applies for the efficiency of A, B, and AB
classes as well with the corresponding conduction angles.
10.2.5 Comparison Between Class A, Class B, Class AB, and Class C Amplifiers
One of the most important reasons for PA’s classification is the difference in the
efficiency and power capability of the classes. Figure 10.30 presents a comparison
between A, B, C, and AB classes based on their conduction angles. According to
Equation 10.51, the efficiency approaches 100% as θ (half the conduction angle, α)
tends to zero, although no power is transferred to the output, for zero conduction
angle. In the class A amplifier, θ will reach its maximum value (180◦ ), leading to an
efficiency of 50%. To modify and control the conduction angle, the current amplitude
can be kept constant while the bias current or the quiescent point is changed. A similar
method is performed by acting the other way around. In both cases, the conduction
angle can be altered as desired.
Power
η capability
id or ic
Class A
α=2π
ωt 100% 0.134
π 2π 3π 4π
0.125
id or ic
Class AB
π<α<2π 78.5%
ωt
π 2π 3π 4π
id or ic
Class B
α=π
ωt
π 2π 3π 4π 50%
id or ic
Class C C B AB A C B AB A
α<π
ωt α α
π 2π 3π 4π 0 π 2π 0 π 2π
1.36π
Figure 10.30: The output current waveforms alongside the efficiencies and the
power capabilities corresponding to different classes of power amplifiers as a
function of the conduction angle, α.
A
L C Vin+ VDC Vin- VDC
Vd1 2:1 Vout 0 0
+ id1 ωt ωt
Vin+
M1 VDD Vd1 2VDD Vd2 2VDD
- RL
Vin- - 0 0
M2 ωt ωt
i
+ d2
Vd2 id1 id2
ωt ωt
π 2π 3π 4π π 2π 3π 4π
Figure 10.31: A typical class D power amplifier topology and the corresponding
voltage and current waveforms.
providing the positive half-cycle transformer current. In the negative half-cycle, the
lower transistor is turned on and the negative half-cycle current flows through the
transformer. Finally, the currents are superimposed to make up the appropriate full-
cycle output signal. A band-pass RLC circuit with high Q suppresses all the higher-
order voltage harmonics at the output. The maximum possible efficiency in class
D amplifiers is 100% and the maximum power capability equals (1/π)Vdmax .Idmax ≈
0.32Vdmax .Idmax . This configuration has a better performance at frequencies far lower
than the unity gain frequency of the transistors, as the transistors are better switches
in this range. Figure 10.31 depicts the corresponding waveforms of class D amplifier.
One of the drawbacks of class D PA’s is that it is not applicable for linear modulations
such as AM, SSB, or DSB in normal circumstances.
As Figure 10.31 shows, the drain voltage of the transistors can maximally swing
up to twice the voltage source value. Assuming that the transistors act as ideal switches,
the voltage at node A can be considered as a rectangular wave swinging symmetrically
between ±VDD . Writing the Fourier series expansion of the voltage at node A, we have
450 Chapter 10. Power Amplifier
The band-pass circuit at the output selects only the main harmonic of vA . Consequently,
assuming that the amplitude of the first harmonic is 4VDD /π at node A, the output
power will be
2
4VDD
VO 2 π 8VDD 2
PO = = = 2 (10.54)
2RL 2RL π RL
In addition, in order to obtain the average current of each transistor in a period,
reminding that the output current is the superposition of the currents of each transistor
at consecutive half-cycles, one can write
T
1 4VDD 4VDD
Z
2
Iave = sin (ωt) dt = 2 (10.55)
T 0 RL π π RL
Ultimately, due to the presence of two active devices, the total power consumption will
be calculated
4VDD 8VDD 2
PDC = 2 ×VDD 2
= 2 (10.56)
π RL π RL
which has exactly the same expression as Equation 10.54. Hence, in the class
D amplifier, the efficiency is ideally 100%. The above relations are based on the
assumption that the switches are ideal while they might have a limited on-resistance
which decreases the efficiency of the PA. If this resistance is not ignored, the output
power of Equation 10.54 will change as follows
2
8VDD 2
RL
PO = 2 (10.57)
π RL ron + RL
The total power consumption in this case will be calculated as
8VDD 2 RL
PDC = × (10.58)
π 2 RL ron + RL
Finally, the maximum efficiency of the PA considering the on-resistance will be
RL
ηmax = (10.59)
ron + RL
The on-resistance might cause the device to have a saturation voltage when turned
on, noted by Vsat . The drain voltage of neither of the devices will thus reach to the
ground level, as they enter the saturation region. Therefore, VDD should be substituted
by VDD −Vsat , and the output power equation will change as
8(VDD −Vsat )2
PO = (10.60)
π 2 RL
10.2 PA Topologies 451
Finally, the maximum efficiency of the PA considering the transistor saturation voltage
will be
VDD −Vsat
ηmax = (10.62)
VDD
In case where the output switches have both a turn-on resistance and a saturation
voltage, the maximum efficiency of a class D amplifier can be expressed as
VDD −Vsat RL
ηmax = × (10.63)
VDD ron + RL
VDD VDD
Vin+ VDC Vin- VDC
RFC RFC 0 0
ωt ωt
Vd1 Vd2
ωt ωt
id1 id2
Vin+ Vin- id1 id2
M1 M2
ωt ωt
π 2π 3π 4π π 2π 3π 4π
Figure 10.32: A typical circuit topology and current waveforms of a class 1/D
power amplifier.
452 Chapter 10. Power Amplifier
Figure 10.33: A typical class E power amplifier topology and the corresponding
voltage and current waveforms.
to the load. Figure 10.33 shows a class E power amplifier. The device voltage and
current waveforms are also demonstrated. As shown in Figure 10.33, the drain voltage
and its first derivative will be zero when the device is turned on. This configuration
has the ability to provide 100% efficiency while its power capability is approximately
0.098Vdmax .Idmax .
In the design of class E power amplifiers, the transistor parasitic capacitances are
crucial due to the frequency limit they impose on the circuit. In the design of the
matching network, the following relations can be used [9]
QRL
L= (10.64a)
ω
1 1
C1 = ≈ (10.64b)
ωRL ((π 2 /4) + 1)(π/2) 5.447ωRL
5.447 1.42
C2 ≈ C1 1+ (10.64c)
Q Q − 2.08
In addition, the maximum output power will be calculated by
VDD 2
Po,max ≈ 0.577 (10.65)
RL
VDD Vin 0v
Tuned to 3f0
L3
RFC ωt
C∞ Tuned to f0 Vout
Vd Vd
C3
Vin
ωt
L1 C1 RL
id
ωt
π 2π 3π 4π
Figure 10.34: A typical class F power amplifier topology and the corresponding
voltage and current waveforms.
9''
5)&
& " Ȝ 7XQHGWRI 9RXW
9G
9LQ
7XQHGWRRGG / & 5/
KDUPRQLFVRII
As observed in section 10.2.6, the active devices act as a switch in class D amplifiers.
The closer the device performance to an ideal switch, the better the resulting efficiency.
One of the methods to ameliorate the switching performance of the device is to increase
the applied voltage to the gate (VGS ). It helps reduce the on-resistance of the device,
leading to increased efficiency. Figure 10.36 shows a general schematic of a class S
power amplifier. A signal converter turns the sinusoidal input signal into a pulse width
modulation (PWM) signal. The PWM signal is applied to the active device in a class
D amplifier, improving its switching performance. Finally, the output signal passes
through a high Q band-pass filter which converts the PWM signal into a sinusoid
(higher-order harmonics are suppressed).
Figure 10.37 shows a typical signal converter. A comparator compares the input
signal to a triangular wave signal which renders a PWM output.
W W W W
Figure 10.36: Block diagram of a class S power amplifier and its corresponding
waveforms.
Modulating
signal
t +
t
- PWM
t signal
Chopping
signal
10.3.1 Back-Off
Nonlinearity clearly reduces the gain and increases the IM at the maximum output
power. So, the simplest way to linearize the PA operation is to force a back-off from
its maximum power. The required back-off depends on the distortion caused by AM
to AM and AM to PM conversions. This method benefits from low cost and no extra
complexity but requires a device with higher power rating (a bigger device). Back-off
Correction Correction
phenomenon is shown in Figure 10.39. The drawback of back-off is that the power
efficiency drops significantly as the back-off is increased.
10.3.2 Predistortion
We can linearize a nonlinear system, knowing its nonlinear transfer function, by
applying the input signal to a system, with an amplitude proportional to the inverse
characteristics of the nonlinear system amplitude response. Ideally, the gain is expected
to remain constant and the phase would vary linearly with frequency (the group
delay would remain constant), and the total system is supposed to remain linear.
Figure 10.39: A typical output power versus input power curve in decibels de-
picting the input back-off and the output back-off with respect to the saturation
point.
10.3 Linearization Techniques in Power Amplifiers 457
Predistortion procedure is shown in Figure 10.40. Let F (u) = G−1 (ku), therefore
−1
y = G G (ku) = ku.
A predistorter conveys the input analog values to the corresponding predistorted
values. A look-up table is perfectly suited for open-loop predistortion. It must be noted
that correcting the distortion caused by the fabrication process, temperature variations,
and aging is difficult. Therefore, it is constructive to upgrade this method to adaptive
predistortion (Figure 10.41). In this method, a demodulator reads the output signal,
followed by an A/D which digitizes it. In the next step, the adaptor block updates the
look-up table data by comparing the input and output signals which have been read.
Finally, a digital predistorter applies the required predistortion to the signal using
the look-up table data. The adaptor loop keeps working until the output signal is
fully corrected. It can be stated that the distortion owing to AM to AM or AM to PM
conversions can be compensated via applying an adaptive predistorter. The drawbacks
of this method include the adaptor loop linearity and delay flatness, and the problems
the look-up table preparation and updating may impose.
Predistorter PA
F(u) G(u)
x F(x) y=G(F(x))=Kx
'LJLWDO /2
1RQOLQHDU
SUHGLVWRULRQ 3$
6RXUFH
VLJQDO 'LJLWDOWR
4XDGUDWXUH
DQDORJ
PRGXODWRU
FRQYHUWHU
$QDORJWR
4XDGUDWXUH
$GDSWDWLRQ GLJLWDO
GHPRGXODWRU
FRQYHUWHU
/2
θ
I
Figure 10.42: Amplitude and phase representation of a signal in the I–Q plane.
/2 /3)
(QYHORSH
GHWHFWRU
6HULDOWR U
'DWD
SDUDOOHO
FRQYHUWHU $PSOLWXGH
FRQYHUWHU
GDWD
, /3)
4XDGUDWXUH
3$
PRGXODWRU
4 /3)
3KDVH
$GMXVWPHQW
/2
4XDGUDWXUH
GHPRGXODWRU
0DLQ
SDWK *DLQ $
3$ 'HOD\
9LQ 9RXW
$ /LQHDU
DPSOLILHU
'HOD\ $
)HHGIRUZDUG
SDWK
The main advantage of this method is its superior stability compared to the pre-
vious ones. However, it requires a supplementary linear amplifier, adding to design
complexity even though it is dealing with a small-signal amplitude. Furthermore, delay
lines should both provide matching and low loss. Notice that the power amplifier’s
gain and the coupler plus attenuator’s path loss are necessitated to be equal with high
precision. High sensitivity to parameters such as fabrication process, temperature, and
aging can be mentioned as this system’s drawback.
10.3.6 Linear amplification with nonlinear components
Linear amplification with nonlinear components is based on converting the amplitude
information into phase and amplifying constant-envelope signals. This method provides
capability of linear amplification at high output powers and it is shown in Figure 10.46.
Since the efficiency is proportional to the average output power, this configuration
is not efficient for modulations with large peak to average ratio (PAR). To improve
linearity, envelope feedback can be used. In the topology depicted in Figure 10.46,
6W
3$
6W
6W 6LJQDO .6W ĮW
6W
VHSDUDWRU ĮW
6W
3$
6W
The amplitude information is hence placed within the signal phase at the output signal
of the separator and the signal envelope will take a constant value, leading nonlinearity
to have a less adverse impact on the data. The total output signal would have the
following form
sout (t) = 2AK cos(α(t)) cos(ωt + φ (t)) = Kb(t) cos(ωt + φ (t)) (10.69)
One of the drawbacks of this method is the analog implementation of the input signal
separator which could be demanding.
(QYHORSH
DPSOLILHU
(QYHORSH 3RZHU
$
GHWHFWRU FRQGLWLRQHU
3RZHU
VXSSO\
(QYHORSH YJ
9''
FRQWURO %LDV
FRQWURO 97+
5)&
& 9RXW
&RQVWDQW 9G
& LG W
HQYHORSH
5)LQ YJ 5/
'ULYHUVWDJH
W
As shown in Figure 10.48, the signal with constant envelope and the bias voltage
have been both applied to the transistor gate. The voltage bias value is calculated
through an envelope control circuit. By applying this input signal, square sine-wave-
tips currents are generated in the drain that can reduce the device loss and thus increase
its efficiency. The impedance matching network is placed at the output. In the design of
the bias circuit, it must be noted that the bandwidth must be higher than the envelope’s
maximum frequency. The efficiency of this architecture is higher than class A and
class AB, and lower than class C. Applying the envelope feedback can result in further
improvements in linearity.
9''
[
/'
Ɛ Ȝ
'ULYHU =
9''
[
/' & 7XQHGWRI 9RXW
Ɛ Ȝ
9LQ
'ULYHU =
5/
9''
[
/'
Ɛ Ȝ
'ULYHU =
Figure 10.49: Block diagram of a switchable parallel amplifier array used for
power control.
Each PA has been designed for a certain value of output power. The output of these
PAs is connected to a transmission line with characteristic impedance, Z0,i with λ /4
length which isolates either of them once its switch is short-circuited. The amplifier
array is connected to the load by means of a 3-bit switch control. The weighted
combination of the amplifiers’ powers appears at the load depending on the switch
states [11]. As such, there would be seven (23 − 1) levels of controllable power which
can be interpreted as a linearity improvement as a function of the input. Furthermore,
the efficiency would be ameliorated as well. Switches must have low loss to guarantee
desirable performance. This architecture has a moderate design complexity.
10.4 Conclusion
In this chapter, we discussed the general requirements and characteristics of the power
amplifiers. Normally, the power amplifier is a block which generates harmonics as
well as IM products due to its nonlinear operation. Furthermore, its gain is normally
saturated by increasing the input amplitude. AM to AM conversion and AM to PM
conversion are among the most important nonlinear characteristics of a power amplifier.
Power amplifiers have different topologies and modes of operation. Class A, class B,
class AB, class C, class D, class E, class F, and class S power amplifiers were introduced
in this chapter along with their corresponding efficiencies and power capabilities.
Linearization techniques in power amplifiers are of great importance in the mod-
ern RF circuitry. Different linearization techniques were introduced in this chapter
including predistortion, polar modulation feedback, Cartesian modulation feedback,
464 Chapter 10. Power Amplifier
10.6 Problems
Problem 10.1 In a handheld transceiver as shown in Figure 10.50, the power am-
plifier has a maximum RF output power of 20 dBm, the transmitter bandpass filter
has a 2 dB insertion loss, and the duplexer’s isolation is about 30 dB. If the receiver
has a bandpass filter with 1 dB insertion loss and 90 dB out-of-band rejection at fTX ,
determine the minimum possible sensitivity of the receiver for a required C/I of 8 dB.
1dB
Duplexer LNA
PA
2dB 20dBm
i = αv + β v2 + δ v5 (10.70)
For an input with the following form, determine the third-order IM products, further-
more, show that if the modulating signals a (t) and b (t) are band limited to W , the
bandwidth of the third-order IM product components would be 5 W:
RS Vout
+
vS Vgs Cgs(V) gmVgs rds Cds RL
-
RS=50Ω rds=500Ω
C0=200fF Cds=90fF
V0=1V RL=500Ω
gm=4mS
Figure 10.51: The equivalent circuit of the MOS stage with a nonlinear gate–
source capacitance.
Problem 10.4 A power MOSFET has the following square-law transfer characteris-
tics IDS = k (VGS −VTH )2 , with a threshold voltage of 0.2V , and k = 50 mA
V2
. We intend
to design a class A power amplifier with VDD = 1V and RL = 100 Ω. First, determine
the required bias point of the transistor, and the maximum drain current and the maxi-
mum drain voltage. Secondly, determine the gain of the stage, and the required input
voltage swing to achieve the maximum output power. What would be the maximum
output power in this case?
where VGS0 ≤ VTH . Determine the output current conduction angle as a function of
V1 and VGS0 , and the output current waveform in this case. Now, compute the DC
component and the first harmonic component of the output current, and consequently,
compute the AC and the DC powers and the efficiency of this amplifier for class B and
class C operation.
10.6 Problems 467
VG<VTH VDD
RFC LD
CD Vout
RS C∞
Vin
RL
VS
Input matching
network
Figure 10.52: A MOSFET class B or class C power amplifier with bias and
matching circuits.
Index
469
470 INDEX
Forouhar Farzaneh was born in Tehran, Iran in 1957. He received his B.S. in
Electrical Engineering from the University of Shiraz, in 1980, Master degree
from E.N.S.T., Paris in1981, DEA and Doctorate from University of Limoges,
France in 1982 and 1985, respectively. He was with Tehran Polytechnic from
1985 to 1989. Since 1989 he has been with the Department of Electrical
Engineering, Sharif University of Technology where he is a Professor. He was
the Chairman of the Department of Electrical Engineering, from 1992 till 1995.
His main areas of interest are Nonlinear RF Circuits, Microwave and
Millimeter wave systems, Antenna Arrays and Wireless Communications. He
is the author of a book in the field of Communication Circuits in Persian,
published by Sharif University Press.
He has been a Senior Member of IEEE since 1997. He was a co-recipient
of the Microwave Prize-European Microwave Conference in 1985, a recipient
of the Maxwell Premium of IEE, U.K. in 2001, and co-recipient Mojtahedi
Innovation Award (Sharif University of Technology) in 2010. He was also
the recipient of the 2015 Hakkak award, in recognition of his tremendous
life-time contribution to national development and propagation of research in
Communications Engineering presented by IEEE-Iran Section.
Ali Fotowat was born in Tehran, Iran, in 1958. He received the B.S. degree in
Electrical Engineering from the California Institute of Technology, Pasadena,
CA, USA, in 1980, and the M.S. and Ph.D. degrees in Electrical Engineering
from Stanford University, Stanford, CA, USA, in 1982 and 1991, respectively.
492 Authors Biographies
Ali Nikoofard was born in Tehran, Iran, in 1990. He received his B.S. in
Electrical Engineering from Shahed University, Tehran, in 2012, M.S. in Elec-
tronics from Sharif University of Technology, Tehran, in 2014, and M.S. from
Case Western Reserve University, Cleveland, OH, USA, in 2017 in integrated
circuit design. He was with KavoshCom Asia R&D Company from 2014 to
2015. He is now working toward Ph.D. at University of California at San Diego,
La Jolla, CA, USA, on ultra-low-power transceiver design with new power
efficient modulation schemes. He is a member of the IEEE Solid-State Circuits
Society since 2013. His current research interests include wireless transceivers,
frequency synthesizers, phase-locked loops and ultra-low-power circuit.
Authors Biographies 493
Mohammad Elmi was born in Tehran, Iran, in 1988. He received his B.S.
degree in Electrical Engineering from Noshirvani University of Technology,
Babol, Iran, in 2013, and M.S. degree in Electrical Engineering (Analog Elec-
tronics) from Shahid Beheshti University, Tehran, Iran, in 2015. He has been
with KavoshCom Asia R&D Company since 2015. He is now working on a
low-power wireless heart monitoring system as a member of research team at
KavoshCom Asia. His research interests include RF integrated circuits design,
low-power analog circuit design, wireless communication transceivers, and
mm-wave integrated circuits.
River Publishers Series in Circuits and Systems
Introduction to Wireless
Communication Circuits
Introduction to Wireless
Communication Circuits
Introduction to Wireless
2nd Edition
2nd Edition