Download as pdf or txt
Download as pdf or txt
You are on page 1of 114

1Phase Noise and I quenry Sithil@ In0MIIlami

-
a - - --.--
- - 8
- --
'I
E l m
5 3-L
rU .= .LO
555
GE,.
s Lo -*
W
g
L!
S S . 2
u c W
5
s!.? J ?E! s
=
;;;
o c v l
5
n
$ 4 g
.- .$ r i z
E a. ' mr t;
u a g 2
% 5.E
.c
G L 5 $
jsz*6
@ a s u p
ru%rr,, s 8m w
w 0
c r
g.2 2 w w
,- L 0 M-0
rn- o
E =,2 $ E
Y O,
X kJ x 3 0
,.-' Z,"m c a
L E P 1 ° 2
: w 0 3 =
0 w z . 2 0
U r d U > U
The Cambridge RF and Microwave Engineering Series
Phase No se and Frequency
Series Editor
Steve C. Cripps

Peter Aaen, Jaime Pla and John Wood, Modeling and Characterization of RF and
Microwave Power FETs
Enrico Rubiola, Phase Noise and Freqziency Stability in Oscillators ENRICO RUBIOLA
Dominique Schreurs, Mairtin O'Droma, Anthony A. Goacher and Michael Gadringer, Professor of Electronics
RF A~~zpliJier
Belzavioral Modeling FEMTO-ST Institute
Fan Yang and Yahya Rahmat-Samii, Electro~nagneticBand Gap Strfluctziresin Antenna CNRS and Universite de Franche Comte
Engineering Besan~on,France

Forthcoming:
Sorin Voinigescu and Timothy Dicltson, High-Frequency Integrated Circuits
Debabani Choudhury, Millimeter Wavesfor Conznzercial Applications
J. Stephenson Kenney, RF Power Anzpl$er Desigrz and Linearization
David B. Leeson, Microwave Sj~stemsand Engineering
Stepan Lucyszyn, Advanced RF MEMS
Earl McCune, Practical Digital Wir)elessConz~zzzinicationsSignals
Allen Podell and Sudipto Chalaaborty, Practical Radio Design Techniques
Patrick Roblin, Nonlinear RF Circziits and the Large-Signal Networlz Analyzer
Dominique Schreurs, Microwave Techniquesfor Microelectronics
John L. B. Wallter, Eiandboolz of RF and Microwave Solid-State Power AnzpliJiers

CAMBRIDGE
UNIVERSITY PRESS
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New Yorlc, Melbourne, Madrid, Cape Town, Singapore, S5o Paulo, Delhi Contents
Cambridge University Press
The Edinburgh Building, Cambridge CB2 8RU, UI<

Published in the United States of America by Cambridge University Press, New York

www.cambridge.org
Information on this title: www.cambridge.org/9780521886772

O Cambridge University Press 2009

This publication is in copyright. Subject to statutory exception


and to the provisions of relevant collective licensing agreements, Foreword by Lute Malelti page ix
no reproduction of any part may take place without Foreword by David Leeson xii
the written permission of Cambridge University Press. xv
Preface
First published 2009 How to use this book xvi
Supplementary material xviii
Printed in the United Kingdom at the University Press, Cambridge Notation xix
A catalog recorzlfor tliispublication is availablefi.onl the Britislz Lib1m.j~
Phase noise and frequency stabilib
ISBN 978-0-52 1-88677-2 hardback
1.1 Narrow-band signals
1.2 Physical quantities of interest
1.3 Elements of statistics
1.4 The measurement of power spectra
1.5 Linear and time-invariant (LTI) systems
1.6 Close-in noise spectrum
1.7 Time-domain variances
1.8 Relationship between spectra and variances
1.9 Experimental techniques
Exercises

Phase noise in semiconductors and amplEiers


2.1 Fundamental noise phenomena
2.2 Noise temperature and noise figure
2.3 Phase noise and amplitude noise
2.4 Phase noise in cascaded amplifiers
2.5 + Low-flicker amplifiers
2.6 + Detection of microwave-modulated light
Exercises

- ---
Heuristic approach to the Leeson effect
Cambridge University Press has no responsibility for the persistence or
accuracy of URLs for external or third-party internet websites referred to
3.1 Oscillator fundamentals
in this publication, and does not guarantee that any content on such 3.2 The Leeson formula
websites is, or will remain, accurate or appropriate.
Contents
Contents vi i

3.3 The phase-noise spectruni of real oscillators 6.7 Optoelectronic oscillators


3.4 Other types of oscillator NIST 10 GHz opto-electronic oscillator (OEO)
OEwaves Tidalwave (10 GHz OEO)
Phase noise and feedback theory Exercises

Resonator differential equation Appendix A Laplace transforms


Resonator Laplace transform
The oscillator References
Resonator in phase space Index
Proof of the Leeson formula
Frequency-fluctuation spectrum and Allan variance
a* A different, more general, derivation of the resonator
phase response
a+ Frequency transformations

Noise in delay-line oscillators and lasers


Basic delay-line oscillator
Optical resonators
Mode selection
The use of a resonator as a selection filter
Phase-noise response
Phase noise in lasers
Close-in noise spectra and Allan variance
Examples

Oscillator hacking
6.1 General guidelines
6.2 About the examples of phase-noise spectra
6.3 Understanding the quartz oscillator
6.4 Quartz oscillators
Oscilloquartz OCXO 8600 (5 MHz AT-cut BVA)
Oscilloquartz OCXO 8607 (5 MHz SC-cut BVA)
RAKON PHARAO 5 MHz quartz oscillator
FEMTO-ST LD-cut quartz oscillator (10 MHz)
Agilent 10811 quartz (10 MHz)
Agilent noise-degeneration oscillator (10 MHz)
Wenzel501-04623 (100 MHz SC-cut quartz)
6.5 The origin of instability in quartz oscillators
6.6 Microwave oscillators
Miteq DRO mod. D-210B
Poseidon DRO- 10.4-FR (1 0.4 GHz)
Poseidon Shoebox (10 GHz sapphire resonator)
UWA liquid-N whispering-gallery 9 GHz oscillator
Phase Noise and Frequency Stability in Oscillators

Presenting a comprehensive account of oscillator phase noise and frequency stability,


this practical text is both mathematically rigorous and accessible. An in-depth treatment
of the noise mechanism is given, describing the oscillator as a physical system, and
showing that simple general laws govern the stability of a large variety of oscillators
differing in technology and frequency range. Inevitably, special attention is given to am-
plifiers, resonators, delay lines, feedback, and fliclter (110noise. The reverse engineering
of oscillators based on phase-noise spectra is also covered, and end-of-chapter exercises
are given. Uniquely, numerous practical examples are presented, including case studies
talten from laboratory prototypes and commercial oscillators, which allow the oscillator
internal design to be understood by analyzing its phase-noise spectrum. Based on tuto-
rials given by the author at the Jet Propulsion Laboratory, international IEEE meetings,
and in industry, this is a useful reference for academic researchers, industry practitioners,
and graduate students in RF engineering and communications engineering.

Additional materials are available via www.cambridge.org/rubiola.

Enrico Rubiola is a Senior Scientist at the CNRS FEMTO-ST Institute and a Professor
at the Universiti: de Franche Comti:. With previous positions as a Professor at the
Universite Henri Poincari:, Nancy, and in Italy at the University Parma and the
Politecnico di Torino, he has also consulted at the NASAICaltech Jet Propulsion
Laboratory. His research interests include low-noise oscillators, phaselfrequency-noise
metrology, frequency synthesis, atomic frequency standards, radio-navigation systems,
precision electronics from dc to microwaves, optics and gravitation.
Foreword by Lute Ma

Given the ubiquity of periodic phenomena in nature, it is not surprising that oscillators
play such a fundamental role in sciences and technology. In physics, oscillators are the
basis for the understanding of a wide range of concepts spanning field theory and linear
and nonlinear dynamics. In technology, oscillators are the source of operation in every
communications system, in sensors and in radar, to name a few. As man's study of
nature's laws and human-made phenomena expands, oscillators have found applications
in new realms.
Oscillators and their interaction with each other, usually as phase locking, and with
the environment, as manifested by a change in their operational parameters, form the
basis of our understanding of a myriad phenomena in biology, chemistry, and even
sociology and climatology. It is very difficult to account for every application in which
the oscillator plays a role, either as an element that supports understanding or insight or
an entity that allows a given application.
In all these fields, what is important is to understand how the physical parameters
of the oscillator, i.e. its phase, frequency, and amplitude, are affected, either by the
properties of its internal components or by interaction with the environment in which
the oscillator resides. The study of oscillator noise is fundamental to understanding all
phenomena in which the oscillator model is used in optimization of the performance of
systems requiring an oscillator.
Simply stated, noise is the unwanted part of the oscillator signal and is unavoidable
in practical systems. Beyond the influence of the environment, and the non-ideality of
the physical elements that comprise the oscillator, the fundamental quantum nature of
electrons and photons sets the limit to what may be achieved in the spectral purity of the
generated signal. This sets the fundamental limit to the best performance that a practical
oscillator can produce, and it is remarltable that advanced oscillators can reach it.
The practitioners who strive to advance the field of oscillators in time-and-frequency
applications cannot be content with laowledge of physics alone or engineering alone.
The reason is that oscillators and cloclts, whether of the common variety or the advanced
type, are complex "systems" that interact with their environment, sometimes in ways
that are not readily obvious or that are highly nonlinear. Thus the physicist is needed to
identify the underlying phenomenon and the parameters affecting performance, and the
engineer is needed to devise the most effective and practical approach to deal with them.
The present monograph by Professor Enrico Rubiola is unique in the extent to which it
satisfies both the physicist and the engineer. It also serves the need to understand both
x Forewords

the fundamentals and the practice of phase-noise metrology, a required tool in dealing and history. But these could not inform further such a comprehensive and extremely
with noise in oscillators. useful book on the subject of oscillator noise.
Rubiola's approach to the treatment of noise in this boolt is based on the input-
output transfer functions. While other approaches lead to some of the same results, Lute Malelti
this treatment allows the introduction of a mathematical rigor that is easily tractable by NASAICaltech Jet Propulsion Laboratory
anyone with an introductory laowledge of Fourier and Laplace transforms. In particular, and OEwaves, Inc.,
Rubiola uses this approach to obtain a derivation, from first principles, of the Leeson February 2008
formula. This formula has been used in the engineering literature for the noise analysis
of the RF oscillator since its introduction by Leeson in 1966. Leeson evidently arrived
at it without realizing that it was laown earlier in the physics literature in a different
form as the Schawlow-Townes linewidth for the laser oscillator. While a number of
other approaches based on linear and nonlinear models exist for analyzing noise in
an oscillator, the Leeson formula remains particularly useful for modeling the noise
in high-performance oscillators. Given its relation to the Schawlow-Townes formula,
it is not surprising that the Leeson model is so useful for analyzing the noise in the
optoelectronic oscillator, a newcomer to the realm of high-performance microwave and
millimeter-wave oscillators, which are also treated in this boolt.
Starting in the Spring of 2004, Professor Rubiola began a series of limited-time
tenures in the Quantum Sciences and Technologies group at the Jet Propulsion Labo-
ratory. Evidently, this can be regarded as the time when the initial seed for this book
was conceived. During these visits, Rubiola was to help architect a system for the
measurement of the noise of a high-performance microwave oscillator, with the same
experimental care that he had previously applied and published for the RF oscillators.
Characteristically, Rubiola had to laow all the details about the oscillator, its principle
of operation, and the sources of noise in its every component. It was only then that he
could implement the improvement needed on the existing measurement system, which
was based on the use of a long fiber delay in a homodyne setup.
Since Rubiola is an avid admirer of the Leeson model, he was interested in applying
it to the optoelectronic oscillator, as well. In doing so, he developed both an approach
for analyzing the performance of a delay-line oscillator and a scheme based on Laplace
transforms to derive the Leeson formula, advancing the original, heuristic, approach.
These two treatments, together with the range of other topics covered, should malte
this unique boolc extremely useful and attractive to both the novice and experienced
practitioners of the field.
It is delightful to see that in writing the monograph, Enrico Rubiola has so openly
bared his professional persona. He pursues the subject with a blatant passion, and
he is characteristically not satisfied with "dumbing down," a concept at odds with
mathematical rigor. Instead, he provides visuals, charts, and tables to malte his treatment
accessible. He also shows his commensurate tendencies as an engineer by providing
numerical examples and details of the principles behind instruments used for noise
metrology. He balances this with the physicist in him that loolts behind the obvious for
the fundamental causation. All this is enhanced with his mathematical skill, of which he
always insists, with characteristic modesty, he wished to have more. Other ingredients,
missing in the boolt, that define Enrico Rubiola are his lcnowledge of ancient languages
Forewords xiii

Foreword by David Leeson The first step of the program to craft a standard that would define frequency stability
was to understand and meld the frequency- and time-domain descriptions of phase
instability to a degree that was predictive and permitted analysis and optimization. By
the time the subcommittee edited the Proc. IEEE special issue, the wide exchange of
viewpoints and concepts made it possible to synthesize concise summaries of the work
in both domains, of which my own model was one.
The committee published its "Characterization of frequency stability" in IEEE Trans.
Instrum. Meas., May 1971. This led to the IEEE 1139 Standards that have served the
community well, with advances and revisions continuing since their initial publication.
Rubiola9s book, based on his extensive seminar notes, is a capstone tutorial on the
Permit me to place Enrico Rubiola's excellent book Phase Noise and Frequency Stability theoretical basis and experimental measurements of oscillators for which phase noise
in Oscillato~*s in context with the history of the subject over the past five decades, going and frequency stability are primary issues.
back to the beginnings of my own professional interest in oscillator frequency stability. In his first chapter Rubiola introduces the reader to the fundamental statistical de-
Oscillator instabilities are a fundamental concern for systems tasked with keeping and scriptions of oscillator instabilities and discusses their role in the standards. Then in the
distributing precision time or frequency. Also, oscillator phase noise limits the demod- second chapter he provides an exposition of the sources of noise in devices and circuits.
ulated signal-to-noise ratio in cominunication systems that rely on phase modulation, In an instiuctive analysis of cascaded stages, he shows that, for modulative or parametric
such as microwave relay systems, including satellite and deep-space links. Comparably fliclter noise, the effect of cascaded stages is cumulative without regard to stage gain.
important are the dynamic range limits in multisignal systems resulting from the maslt- This is in contrast with the well-lmown treatment of additive noise using the Friis
ing of small signals of interest by oscillator phase noise on adjacent large signals. For formula to calculate an equivalent input noise power representing noise that may originate
example, Doppler radar targets are maslced by ground clutter noise. anywhere in a cascade of real amplifiers. This example highlights the concept that "the
These infrastructure systems have been well served by what might now be termed model is not the actual thing." He also describes concepts for the reduction of fliclter
the classical theory and measurement of oscillator noise, of which this volume is a noise in amplifier stages.
comprehensive and up-to-date tutorial. Rubiola also exposes a number of significant In his third chapter Rubiola then combines the elements of the first two chapters to
concepts that have escaped prior widespread notice. derive models and techniques useful in characterizing phase noise arising in resonator
My early interest in oscillator noise came as solid-state signal sources began to be feedback oscillators, adding mathematical formalism to these in the fourth chapter. In
applied to the radars that had been under development since the days of the MIT Radiation the fifth chapter he extends the reader's view to the case of delay-line oscillators such
Laboratory. I was initiated into the phase-noise requirements of airborne Doppler radar as lasers. In his sixth chapter, Rubiola offers guidance for the instructive "haclting" of
and the underlying arts of crystal oscillators, power amplifiers, and nonlinear-reactance existing oscillators, using their external phase spectra and other measurables to estimate
frequency multipliers. their internal configuration. He details cases in which resonator fluctuations mask circuit
In 1964 an IEEE committee was formed to prepare a standard on frequency stability. noise, showing that separately quantifying resonator noise can be fruitful and that device
Thanlts to a supportive mentor, W. K. Saunders, I became a member of that group, which noise figure and resonator Q are not merely arbitrary fitting factors.
included leaders such as J. A. Barnes and L. S. Cutler. It was noted that the independent It's interesting to consider what lies ahead in this field. The successes of today's
use of frequency-domain and time-domain definitions stood in the way of the develop- consumer wireless products, cellular telephony, WiFi, satellite TV, and GPS, arise directly
ment of a common standard. To promote focused interchange the group sponsored the from the economies of scale of highly integrated circuits. But at the same time this
November 1964 NASAIIEEE Conference on Short Term Frequency Stability and edited introduces compromises for active-device noise and resonator quality. A measure of
the February 1966 Special Issue on Frequency Stability of the Proceedings of the IEEE. the marltet penetration of multi-signal consumer systems such as cellular telephony
The context of that time included the appreciation that self-limiting oscillators and and WiFi is that they attract enough users to become interference-limited, often from
many systems (FM receivers with limiters, for example) are nonlinear in that they subscribers much nearer than a distant base station. Hence low phase noise remains
limit amplitude variations (AM noise); hence the focus on phase noise. The modest essential to preclude an unacceptable decrease of dynamic range, but it must now be
frequency limits of semiconductor devices of that period dictated the common usage of achieved within narrower bounds on the available circuit elements.
nonlinear-reactance frequency multipliers, which multiply phase noise to the point where A search for new understanding and techniques has been spurred by this requirement
it dominates the output noise spectrum. These typical circuit conditions were second for low phase noise in oscillators and synthesizers whose primary character is integration
nature then to the "short-term stability community" but might not come so readily to and its accompanying minimal cost. This body of laowledge is advancing through
mind today. a speculative and developrnental phase. Today, numerical nonlinear circuit analysis
xiv Forewords

supports additional design variables, such as the timing of the current pulse in nonlinear
oscillators, that have become feasible because of the improved capabilities of both Preface
semiconductor devices and computers.
The field is alive and well, with emerging players eager to find a role on the stage for
their own scenarios. Professionals and students, whether senior or new to the field so ably
described by Rubiola, will benefit from his theoretical rigor, experimental viewpoint,
and presentation.

David B. Leeson
Stanford University
February 2008 The importance of oscillators in science and technology can be outlined by two mile-
stones. The pendz~lum,discovered by Galileo Galilei in the sixteenth century, persisted
as "the" time-measurement instrument (in conjunction with the Earth's rotation period)
until the piezoelectric quartz resonator. Then, it was not by chance that the first inte-
grated cir~z~it,built in September 1958 by Jack Kilby at the Bell Laboratories, was a
radio-frequency oscillator.
Time, and equivalently frequency, is the most precisely measured physical quantity.
The wrist watch, for example, is probably the only cheap artifact whose accuracy ex-
ceeds 1o - ~ ,while in primary laboratories frequency attains the incredible accuracy of
a few parts in 10-15. It is therefore inevitable that virtually all domains of engineering
and physics rely on time-and-frequency metrology and thus need reference oscillators.
Oscillators are of major importance in a number of applications such as wireless com-
munications, high-speed digital electronics, radars, and space research. An oscillator's
random fluctuations, referred to as noise, can be decomposed into amplitude noise and
phase noise. The latter, far more important, is related to the precision and accuracy of
time-and-frequency measurements, and is of course a limiting factor in applications.
The main fact underlying this book is that an oscillator turns the phase noise of its
internal parts into frequency noise. This is a necessary consequence of the Barlchausen
condition for stationary oscillation, which states that the loop gain of a feedback oscillator
must be unity, with zero phase. It follows that the phase noise, which is the integral of
the frequency noise, diverges in the long run. This phenomenon is often referred to as
the "Leeson model" after a short article published in 1966 by David B. Leeson [63]. On
my part, I prefer the term Leeson effect in order to emphasize that the phenomenon is
far more general than a simple model. In 2001, in Seattle, Leeson received the W. G.
Cady award of the IEEE International Frequency Control Symposium "for clear physical
insight and [a] model of the effects of noise on oscillators."
In spring 2004 I had the opportunity to give some informal seminars on noise in oscil-
lators at the NASAICaltech Jet Propulsion Laboratory. Since then I have given lectures
and seminars on noise in industrial contexts, at IEEE symposia, and in universities and
government laboratories. The purpose of most of these seminars was to provide a ttlto-
rial, as opposed to a report on advanced science, addressed to a large-variance audience
that included technicians, engineers, Ph.D. students, and senior scientists. Of course,
capturing the attention of such a varied audience was a challenging task. The stimu-
lating discussions that followed the seminars convinced me I should write a working
xv i Preface

document1 as a preliminary step and then this book. In writing, I have made a serious lecturer experimentalist
effort to address the same broad audience.
This work could not have been written without the help of many people. The gratitude 1
at least a quick look,
I owe to my colleagues and friends who contributed to the rise of the ideas contained depending on your
background & need
in this book is disproportionate to its small size: Rimi Brendel, Giorgio Brida, G. John
Dick, Michele Elia, Patrice Firon, Serge Galliou, Vincent Giordano, Charles A. (Chuck)
& electron. dev.
Greenhall, Jacques Groslambert, John L. Hall, Vladimir S. (Vlad) Ilchenko, Laurent
Larger, Lutfallah (Lute) Malelci, Andrey B. Matslto, Marlt Oxborrow, Stefania Romisch,
Anatoliy B. Savchenkov, Frangois Vernotte, Nan Yu.
of real oscillators
Among them, I owe special thanks to the following: Lute Malelti for giving me the
opportunity of spending four long periods at the NASAICaltech Jet Propulsion Labora-
tory, where I worlced on noise in photonic oscillators, and for numerous discussions and
suggestions; G. John Dick, for giving invaluable ideas and suggestions during numerous 5 optional I 2 t optional
and stimulating discussions; Rimi Brendel, Marlc Oxborrow, and Stefania Romisch for
delay-line oscill.
their personal efforts in reviewing large parts of the manuscript in meticulous detail and
3 9
(.Ti]
for a wealth of suggestions and criticism; Vincent Giordano for supporting my efforts optional
noise in amplif. phase flicker and
for more than 10 years and for frequent and stimulating discussions. Eiz,er an~p,ers
I wish to thank some manufacturers and their local representatives for kindness and
prompt help: Jean-Pierre Aubry from Oscilloquartz; Vincent Candelier from RAKON
(formerly CMAC); Art Faverio and Charif Nasrallah from Miteq; Jesse H. Searles from goes here
Poseidon Scientific Instruments; and Marlt Henderson from Oewaves.
selected topics
Thanks to my friend Roberto Bergonzo, for the superb picture on the front cover, and examples
entitled "The amethyst stairway." For more information about this artist, visit the website
http://robertobergonzo.com.
Finally, I wish to tha& Julie Lancashire and Sabine ICoch, of the Cambridge editorial
staff, for their kindness and patience during the long process of writing this book.
theoretical practical theoretical practical

How to use this book Figure 1 Asymptotic reading paths: on the left, for someone planning lectures on oscillator
noise; on the right, for someone currently involved in practical work on oscillators.
Let us first abstract this boolt in one paragraph. Chapter 1 introduces the language of
phase noise and frequency stability. Chapter 2 analyzes phase noise in amplifiers, includ-
ing fliclter and other non-white phenomena. Chapter 3 explains heuristically the physical optics; boolcs focusing on the relevant mathematical physics. The present text is unique
mechanism of an oscillator and of its noise. Chapter 4 focuses on the mathematics that in that we look at the oscillator as a system consisting of more or less complex interacting
describe an oscillator and its phase noise. For phase noise, the oscillator turns out to be bloclts. Most topics are innovative, and the overlap with other boolts about oscillators
a linear system. These concepts are extended in Chapter 5 to the delay-line oscillator or time-and-frequency metrology is surprisingly small. This may require an additional
and to the laser, which is a special case of the latter. Finally, Chapter 6 analyzes in effort on the part of readers already familiar with the subject area.
depth a number of oscillators, both laboratory prototypes and commercial products. The The core of this book rises from my experimentalist soul, which later became con-
analysis of an oscillator's phase noise discloses relevant details about the oscillator. vinced of the importance of the mathematics. The material was originally thought and
There are other boolts about oscillators, though not numerous. They can be divided into drafted in the following (dis)order (see Fig. 1): 3 Heuristic approach, 6 Oscillator haclc-
three categories: boolts on radio-frequency and microwave oscillators, which generally ing, 4 Feedback theory, 5 Delay-line oscillators. The final order of subjects aims at a
focus on the electronics; boolts about lasers, which privilege atomic physics and classical more understandable presentation. In seminars, I have often presented the material in the
3-6-4-5 order. Yet, the best reading path depends on the reader. Two paths are suggested
E. Rubiola, The Leeson Efeect - Phase Noise in Quasilirzear. oscillator.^, February 2005, arXiv:physics/ in Fig. 1 for two "asymptotic" reader types, i.e. a lecturer and experimentalist. When
0502143, now superseded by the present text. planning to use this boolt as a supplementary text for a university course, the lecturer
xvi i i Preface

should be aware that students often lack the experience to understand and to appreciate Notation
Chapter 6 (Oscillator hacking) and other practical issues, while the theory can be more
accessible to them. However, some mathematical derivations in Chapters 4 and 5 may
require patience on the part of the experimentalist. The sections marked with one or two
stars, * and **, can be skipped at first reading.

Supplementary material
My web page
http://rubiola.org (also http://rubiola.net) The following notation list is not exhaustive. Some symbols are not listed because
contains material covering various topics about phase noise and amplitude noise. A they are introduced in the main text. On occasion a listed symbol may have a different
section of my home page, at the URL meaning, where there is no risk of ambiguity because the symbol has local scope and
the usage is consistent with the general literature.
http://rubiola.org/oscillator-noise Uppercase is often used for
has been created for the supplementary material specific to this boolc. Oscillator noise Fourier or Laplace transforms
spectra and slides from my seminars are ready. Other material will be added later. constants, when the lower-case symbol is a function of time. For example, in relation
Cambridge University Press has set up a web page for this book at the URL to v(t) we have K.,,, Vo (peak)
www.cambridge.org/rubiola , quantities conventionally represented with an upper-case symbol.
in boldface, phasors. Example, V = ~,,,ej'.
where there is room for supplementary material. It is my intention to malce the same
material available on my home page and on the Cambridge website. Yet, my web page Though w is the arzgttlar jieqtlerzcy, for short it is referred to as the frequency.
is under my full control while the other one is managed by Cambridge University Press. Numerical values are always given in Hz. The symbol w may be used as a shorthand for
2nv or 2nf . The symbols v and f always refer to single-sided spectra and w always
refers to two-sided spectra even if only the positive frequencies appear in plots.
Section 1.2 provides additional information about the relevant physical quantities,
their meaning, and their usage, and about the variables associated with them.
The list irzcltldes some chcqtei~ection,stlbsection, eqziation, orJigtire c ~ ~ oreferences.
ss

Symbol Meaning and text references


A amplifier voltage gain (thus, the power gain is A2)
bi coefficients of the power-law approxiiliation of S,( f ).
1.6.2, (1.70), and Fig. 1.8
resonator phase response. 4.4 and (4.62)
resonator impulse response. 4.7
resonator phase response, B(s) = L{b(t)}. 4.4.2,4.4.3
electrical capacitance, farad
denominator (of a fraction or of a rational function)
energy, either physical (J) or mathematical
(dimensionless), depending on context
electric field, Vlm
mathematical expectation. 1.3.1 and (1.28)
Fourier frequency, Hz. 1.2
generic function. 1.2
a~nplifiercorner frequency, Hz. 2.3.3 equivalent noise temperature of a device. 2.2
Leeson frequency, Hz. 3.2 and (3.2 1) observation or measurement time in truncated signals. 11.4.1
amplifier noise figure. 2.2 and (2.11) period, T = l l v
Fourier transform operator. (A. 3) absolute temperature, reference temperature TO= 290 K
Planck9sconstant, h = 6.626 x Js transmission coefficient. 5.2
coefficients of the power-law approximation of SJ,(f). Heaviside (step) function, U(t) = 6(tf)dt'
1.6.3 and (I .73), (I .74) voltage (in theoretical contexts, also a diniensiorlless signal)
inlpulse response. 1.5.1 and (1.55), (1 3 6 ) a generic varia,ble
phase response pl~ase-timefluctuation. 1.2 and (1. 17)
transfer function, H(s) = L{lz(t)) , also N(jw). fractional-fi-equencyfluctuation. 1.2 and (1.18)
1.5.1 and (4.42) dc or peak voltage
phase transfer fi~nction,H(s) = L{h(t)), also H(jw) Laplace transforan of v(t)
3.2 and 4.5 voltage pliasor. 1.1
current, as a f~~nction of time normalized amplitude noise. 1.1.1
imaginary unit, j = - I transfer function of the feedback path. 4.2 and Fig. 4.6
Boltzmann constant, 1-381 x 1VZ3 J/K Dirac delta function
a constant, lcd9l ~lcL9, etc. difference operator, in Av, (Aw)(t), etc.
harmonic order (in Chapter 5) pliotodetector quantum efficiency. 2.2.3
voltage attenuation or loss (thus, the power loss is t2) phase or argument of a complex function p e j s
electrical inductance, H small phase step. Chapter 4
Laplace transform operator. 1.5.1 and (A. 1) wavelength
single-sideband noise spectrum, dBc/Hz. harmonic order in phase space. Chapter 5
1.tiel and (1 -68) frequency (Hz), used for carriers. I .2
integer (in Chapter 5) modulus of a complex function ,miB
modulation index (of light intensity) photodetector responsivity, AIW. 2.2.3
ra~idoannoise, either near-dc or rf-microwave -+
real part of the coi~plexvariable s = rs j w
integer Allan deviation, square root of the Allan variance a i ( r ) . 1.7
noise power spectral density, W/Hz measurenient time, in oJ,( T )
numerator (of a fraction or of a rational function) resonator relaxation time. 4.1
co~iplexvariable, replaces s when needed delay of a delay line, and group delay of the mode
pov$eq either physical (W) or mathematical selector filter. Chapter 5
(di~nensionless),depending on context phase (constant), phase noise. 1.1
4 electron charge, q = 1.602 x G phase noise, @(s) = L{q(t))
e resonator quality factor. 4.1 dissonance. 4.2 and (4.3 1)
R , 80 resistance, load resistance (often Ro = 50 a) amplifier static phase, phase noise. 3.2 and 4.5
R reflection coefficient. Chapter 5 amplifier phase noise, Q(s) = C{$(t))
R ( r) autocorrelation or correlation function. 1.4.1 angular frequency, carrier or Fourier. 1.1
S complex variable, s = a -tj w oscillator angular frequency. 1.1
S(.f1 power spectral density (PSD). 1.4.1, 1.4.2 Leeson angular frequency
SCI (f> one-sided PSD of the quantity a resonator natural angular frequency. 1.2
Sq (.f) one-sided PSD of the random phase q(t). 1.6.1 resonator free-decay angular pseudo-frequency. 1.2
s'qf) one-sided PSD, 1.4.2. The variable is could also be v replaces w, wlien needed
5'" ( m ) two-sided PSD. (1.4.13. detuning angular frequency. Chapter 5
t time
xxi i Notation

Subscript Meaning
0 oscillator carrier, in coo, Po, Vo, etc.
i input. Examples vi(t), vi(t), cDi(s)
i current. Example, shot noise Si(w) = 2q7
1 light
L Leeson
L loop
m main branch
n resonator natural frequency (a,, v,)
0 output. Examples vo(t), y,,(t), @,(s)
In theoretical physics, the word "oscillator" refers to a physical object or quantity
P resonator free-decay pseudo frequency (q , up)
oscillating sinusoidally - or at least periodically - for a long time, ideally forever,
P +
pole, as in s, = op jwp (referring to a complex variable)
without losing its initial energy. An example of an oscillator is the classical atom, where
P peak. Example, V,, = &KI,,,
the electrons rotate steadily around the nucleus. Conversely, in experinzental science
rms root mean square
the word "oscillator" stands for an artifact that delivers a periodic signal, powered by
z +
zero, as in s, = oz j w, (referring to a complex variable)
a suitable source of energy. In this book we will always be referring to the artifact.
Examples are the hydrogen maser, the magnetron of a microwave oven, and the swing
Symbol Meaning
wheel of a luxury wrist watch. Strictly, a "clock" consists of an oscillator followed by a
<> mean
gearbox that counts the number of cycles and the fraction thereof. In digital electronics,
< >N mean of N values. 1.3.1
- the oscillator that sets the timing of a system is also referred to as the clock. Sometimes
x time average of x, for example. 1.3.1
the term "atomic clock" is improperly used to mean an oscillator stabilized to an atomic
t-) transform-inverse-transform pair. Example, x (t) u X(w)
transition, because this type of oscillator is most often used for timekeeping.
* convolution. Example, vo(t) = h(t) * vi(t). 1.5.1
A large part of this book is about the "pre~ision"~ of the oscillator frequency and
V
A
asymptotically equal
about the mechanisms of frequency and phase fluctuations. Before tackling the main
subject, we have to go through the technical language behind the word "precision," and
present some elementary mathematical tools used to describe the frequency and phase
fluctuations.

1.I Narrow-band signals

The ideal oscillator delivers a signal

v(t) = Vo cos(wot + p) (pure sinusoid) , (1.1)


where Vo = K.,, is the peak amplitude, wo = 2nvo is the angular frequency; and y,
is a constant that we can set to zero. Let us start by reviewing some useful representations
associated with (1.1).
A popular way of representing a noise-free sinusoid v(t) in Cartesian coordinates
is the phasor, also called the F~esnelvector. The phasor is the complex number V =
-+
A j B associated with v(t) after factoring out the wot oscillation. The absolute value

Here, the word "precision" is not yet used as a technical term.


The symbol w is used for the angular frequency. Whenever there is no ambiguity, we will omit the adjective
"angular" and give the numerical value in Hz, which of course refers to vo = w l ( 2 n ) .
Phase noise and frequency stabiliw 1.I Narrow-band signals 3

IVI is equal to the rms value of v(t), and the phase arg V is equal to q . The phase Time donlain Phasor representation
reference is set by cos mot. Alternatively, the phasor is obtained by expanding v(t) as
Vo (cos wet cos q - sin wot sinq). Then, the real part is identified with the rms value
of the cos mot component and the imaginary part with the rms value of the - sin wot
component. Thus, the ideal signal (1.1) may be represented as the phasor

vo eJq.
v=-
2/2 (phasor) .
(1.2) ohase fluctuation
V=
vo
-( c o s q + j s i n q )
2/2
A more powerful tool is the analytic signal z(t) associated with v(t), also called the
pre-envelope and formally defined as

z(t) = v(t) + j G(t) (analytic signal) , (1.3)


Figure 1.1 Amplitude and phase noise: Vo is in volts, a(t) is non-dimensional, p(t) is in radians,
where G(t) is the Hilbert transform of v(t), i.e, v(t) shifted by 90". The analytic signal is and x ( t )is in seconds.
most often used to represent narrowband signals, i.e. signals whose power is clustered
in a narrow band centered at the frequency oo. However, it is not formally required that 1 .I .I The clock signal
the bandwidth be narrow, nor that the power be centered at wo, and not even that wo be
contained in the power bandwidth. Amplitude and phase can be (slowly) time-varying In the real world, an oscillator signal fluctuates in amplitude and phase. We introduce
signals. the quasi-perfect sinusoidal cloclc signal (Fig. I. 1)
The analytic signal z(t) is obtained from v(t ) by deleting the negative-frequency side
of the spectrum and multiplying the positive-frequency side by a factor 2. Alternatively,
z(t) can be obtained using any of the following replacements: The term "cloclt signal" emphasizes the fact that the cycles of v(t), and fractions thereof,
can be counted by suitable circuits, so that v(t) sets a time scale. When talking about
cloclts, we assume that (1.6) has a high signal-to-noise ratio. Hence we note the following.
vo
v(t) = -cos(wot + q) + z(t) = ~(t)ejq(')ejuo' (1.4)
+ a(t)], where a(t)
2/2: The peak amplitude Vo of (1 .I) is replaced by the envelope Vo[l
z(t) = V(t)(cos q + j sin q) ejwot. is the random fractional amplitudes3The assumption

The analytic signal has two relevant properties.

1. A phase shift 6 applied to v(t) is represented as z(t) multiplied by ejQ. reflects the fact that actual oscillators have small amplitude fluctuations. Values la(t)l E
2. Since the power associated with negative frequencies is zero, the total signal power (lo-', are common in electronic oscillators.
can be calculated using the positive frequencies only. The constant phase q of (1.1) is replaced by the random phase q(t), which originates
the clock error. In most cases, we can assume that
The conzplex envelope of z(t), also referred to as the low-pass process associated with
z(t), is obtained by deleting the complex oscillation ejw0' in the analytic signal. The
complex envelope is the natural extension of the phasor and is used when the amplitude
and phase are allowed to vary with time: A slowly varying phase is often referred to as drift. Observing a cloclt in the long
term, the assumption Iq(t)l << 1 is no longer true. Yet it is possible to divide the carrier
frequency by a suitably large rational number. The phase scales down accordingly, so
that the condition Iq(t)l << 1 is obtained at low frequencies.
Strictly speaking, the phasor refers to a pure sinusoid. Yet the terms "phasor" and
"time-varying phasor" are sometimes used in lieu of the term "complex envelope." The symbol ~ ( tis)often used in the literature instead of a(t).
4 Phase noise and frequency stabiliw 1.2 Physical quantities of interest 5

a time T,longer than the oscillation period yet shorter than the time scale T, of the
amplitude fluctuations, is

,, .,, v,'
p(t)=-[1+2a(t)], la(t)l<<l, 2n/wo<<Tm<<T,. (1.14)
(VOID)
cos mot -9:
,a ,,
:<- 2Ro
(vx ID) COS mot
Figure 1.2 Phasor representation of a noisy sinusoid. Example 1.1. Let us estimate the error accumulated in 1 year by a cloclt based
on a 10 MHz oscillator accurate to within lo-''. The maximum cloclt error is
The cloclt signal can be rewritten in Cartesian coordinates by separating the cos mot Te = (Aw/o)T,,,,; thus T, = 10-lo x 3.16 x lo7 s = 3.2 ms in one year. The oscil-
and - sin wot components4 (Fig. 1.2): lation period is T, = 2n/wo = s. Hence the cloclc error accumulated in one year is
n = T,/T, = 3.16 x lo4 cycles ofthe 10 MHz carrier; thus q = 2nn = 2 x lo5 rad.
v(t) = Vo cos wot + v , (t)
~ cos wot - vJ,(t)sin wot . (1.9)
The signals v,(t) and vJ,(t) are called the in-phase and qziud~fiatzi~e
components of the
noise, respectively. Physical quantities of interest
The polar representation (1.6) and the Cartesian representation (1.9) are connected by

/ + 91'+ [$I
Traditionally, physicists use the symbol v for the frequency while electrical engineers
2
prefer f . It is unfortunate that in the domain of time-and-frequency metrology the
a(t) = [1 -1 (fractional amplitude) , (1.10)
notation is sometimes unclear or difficult to understand because both v and f are found
in the same context. When I came to metrology in the early 1980s with a background in
p(t) = arctan vJ,(t)
electronics and telecommunications, it took me a long time to get used to this unnecessary
Vo + V,Y (t) complication. The early articles about frequency stability use v for fixed frequencies,
Equation (1.10) is the Pythagorean theorem written in terms of the real component such as a carrier or a beat note, and f for the Fourier frequency, i.e. the variable of
+
( Vo v,~)/& and the imaginary component vJ,/&. Equation (1.1 1) is arctan 5/%,i.e. spectral analysis. Other articles use v in the carrier signal cos 2nvot and f for the
the arctangent of the imaginary-to-real ratio. A problem with (1.1 1) is that the arctangent spectral analysis of the low-pass fluctuations (a, q, etc.), considering the carrier and
returns the principal value, i.e. the value defined in (-n/2, n/2), or in (-n, n ) using a the low-pass fluctuations as nearly separate worlds. Of course, these two distinctions
two-argument arctangent. The cycles accumulated, if any, are to be counted separately. between v and f are similar, so it may be difficult to decide whether a frequency should
In low-noise conditions, it holds that, for 1 v, / Vo1 << 1 and 1 v , / Vo1 << 1, be v or f . Additional confusion arises from the fact that fluctuations even smaller than
10-l6 are sometimes mea~ured;~ no frequency can be talten as constant. In the end, it is
a(t) = -
vx(t) (fractional amplitude) (1.12) recommended that the exact meaning of a frequency symbol in a particular equation is
Vo always checlted.
Another point is that the oscillator instability can be described as a phase fluctuation
~ ( t= v~'(t)
) ---- (phase) .
Vo or as a frequency fluctuation. The first choice is made in the definition of the cloclt signal
The spectrum of a pure sinusoid such as (1.1) is an ideally thin line at wo, mathe- (1.6), here repeated:
matically described as a Dirac delta function 6(0 - wo). Noise broadens the spectrum:
the cloclt signal (1.6) looks like a line of bandwidth twice that of a(t) and p(t) if the
signal-to-noise ratio is high. In the case of a low signal-to-noise ratio, p(t) yields a In this representation, it is implicitly assumed that coo is the best estimate of the oscillator
linewidth larger than twice the bandwidth. frequency, so that wot describes the oscillation and p(t) describes its phase fluctuation.
Finally, the random phase p(t) does not contribute to the signal power. The instan- This approach is suitable for short-term measurements, where the oscillator stability
taneous power is P(t) = v 2 ( t ) / ~ 0In
. low-noise conditions the power, averaged over is sufficient for p(t) to stay in the interval (-n, n). In the longer term, the oscillator

It is instructive to relate this small value to the internal computer representation of numbers. Since the
The form x ( t ) cos mot - y ( t ) sin mot is preferable for a signal in Cartesian coordinates, but in this chapter IEEE standard "double precision" format has a 15 digit mantissa, the value 10-l6 is one order of magnitude
x and y are used for other relevant quantities. smaller than the roundoff error.
Phase noise and frequency stabiliw 1.2 Physical quantities of interest

ends up drifting more than a half-cycle of the carrier frequency and the phase q(t) oscillator
under test
becomes ambiguous. In such cases, we may prefer to characterize the oscillation using
the frequency fluctuations. Then the clock signal is written as

v(t) = Vo[l + a(t)] cos (clock signal) , (1.15)

where

(Aw)(t) = @(t) (angular-frequency fluctuation) (1.16)

is the angular frequency fluctuation.


Additionally, it is often usehl to normalize q(t) in order to transform the phase noise
into time fluctuations, expressed in seconds (see below), and to normalize the oscillator
frequency.
The definitions summarized below are aimed at giving straightforward access to the
general literature on phase noise and frequency stability; Fig. 1.3 relates the quantities
described to a typical experimental setup.

q(t) represents the phase noise, i.e. the random phase fluctuation defined by (1.6).
a(t) represents the fractional-amplitude noise (for short, "amplitude noise"), i.e. the
random amplitude fluctuation defined by (1.6).
v is used for the carrier frequency, either radio, microwave, or optical, and also for the 3+
3.a+ x
A [fluctuation)
delay)
beat note between two carriers. The symbol v can be a variable, as on the frequency
axis of a spectrum analyzer, or a constant, as in the nearly constant frequency of 1 4 + d / stop T
an oscillator. We can also find v(t) used in the same way as the quantity (Av)(t)
reference
introduced below. oscillator
Example: let vl = 10 GHz and v2 = 10.24 GHz be the frequency of two oscillators.
On the spectrum analyzer, we see two lines at v = vl and v = v2. After mixing, I
I ext
the beat frequency is vb = v2 - vl = 240 MHz. I

'_________--- I ref.
Av = v - vo is the difference between the actual frequency v and a reference value
vo. The latter can be the nominal frequency or a reference value close to v.
Figure 1.3 Simplified diagram of oscillator noise measurements in the time domain, illustrating
(Av)(t) represents the instantaneous frequency fluctuation (or noise). This implies the main physical quantities of interest in time-and-frequency metrology.
the assumption of slow modulation with a high modulation index, so that the signal
can be approximated by a slow-swinging carrier. This means that the carrier and
sidebands degenerate into a single Dirac S hnction that tracks the modulation. Example: inspecting the phase noise of a microwave amplifier in some specific
In most practical cases, (Av)(t) should be regarded as the fluctuating output of a conditions, we find white noise at high f and flicker noise below the corner
frequency comparator, after discarding the dc component. frequency f,.
f is used for frequency in the spectral analysis of low-pass processes, close to dc, w is the angular frequency. The unit associated with w is radls. In this book, w is a
after detection. Thus, f is used in connection with a(t), q(t), v,(t), v,(t), etc. The shorthand for either 2nv or 2nf .We also use A o and (Aw)(t). The word "angular"
symbol f can refer to a variable, as on the frequency axis of a F F T ~analyzer, or is often omitted, and numerical values are given in Hz, which of course refers to
to a constant. w/(2n>.
!2 is used instead of w in some special cases, as in Chapter 5, where we need to
represent several angular frequencies at the same time. Preferentially, !2 refers to
There is no theoretical need to use a FFT (fast Fourier transform) analyzer to measure a near-dc process.
However, this is the type of spectrum analyzer used in virtually all cases. low-pass phenomena.
8 Phase noise and frequency stability

x(t) is the phase-time fluctuation, that is, the random phase fluctuation p(t) converted A time lag can be present from input to output if the synthesizer includes a phase-loclted
into time, and measured in seconds: loop (PLL). However, the fractional frequency fluctuation and the phase-time fluctuation
observed at the output are equal to the input fluctuations:
~ ( t=
) -
cp(t) = -
p(t) (phase-time fluctuation) .
oo 2nv0
Here vo is either the nominal or the estimated frequency.
Interestingly, x(t) does not become ambiguous when p(t) exceeds half a cycle of
the carrier (Fig. 1.3) because it allows the accumulation of phase cycles.
y(t) is the fractional-frequency fluctuation, i.e. the instantaneous frequency fluctua-
tion normalized to the carrier frequency vo. The quantity y(t) is dimensionless. Finally, we notice that there is no general law for the propagatation of the amplitude
fluctuation a(t) through a synthesis chain. The reason is that the synthesis needs strong
nonlinearity, hence the amplitude is saturated.

_---
- - (Av)(t) (fractional-frequency fluctuation) . (1.18)
oo vo Large phase noise
The output of digital instruments is a stream of sainpled values, denoted by an integer The above rules hold when the phase noise is low. In large-phase-noise conditions, the
subscript. Thus, xl, is x(t) sampled at the time t = k z . synthesizer's behavior is governed by the energy conservation law. The easiest way to
Finally, a number of relationships are written in their usual form, found in most understand this is to write the output signal in the analytic form z(t) = K
, ejWde j ~ ( ' ) .
textbooks. Therefore, it is inevitable that some of the above symbols are also used in The phase-noise term ej'(') spreads the power into the noise sidebands, yet without
expressions such as "let f (x) be a function . . . ," etc. changing the total power because lej~(')l= 1. If the output phase noise exceeds some
Three frequencies have a special rGle all through this book and are used extensively 2 radians, the sidebands sink most of the power and the carrier power drops abruptly. This
in Chapters 4 and 5 ; thus they deserve to be mentioned here. phenomenon is referred to as carrier collapse. As a consequence, extremely high spectral
purity is needed when the multiplication ratio is high, for example in the synthesis of
oo and vo are the oscillator frequency and angular frequency, i.e. those of the carrier.
THz or optical signals from electronic oscillators.
onand vn are the natural angular frequency and natural frequency of a resonator.
Mathematically, carrier collapse is a consequence of the application of the Angers-
w, and v, are the free-decay pseudofrequency and angular pseudofrequency of a
Jacobi expansion
resonator. It holds that o, 5 oo.

In most oscillators, the oscillation frequency wo is determined by the natural frequency


w, of a resonator. However, o o differs slightly from o,,because of feedback.

1.2.1 * Frequency synthesis to the angular modulation, according to which the carrier amplitude is dominated by the
In a large number of applications the oscillator is the reference of a frequency synthesizer. --
Bessel function Jo(z). The function Jo(z) nulls at z 2.405. Further consideration of
A noise-free synthesizer can be regarded as a gearbox that multiplies the input frequency this phenomenon is beyond our scope, however.
mi by a rational number N I D and outputs a frequency

Elements of statistics
In the presence of small fluctuations, the input fluctuations propagate to the synthesizer
output with the same N / D law, that is, In the previous sections, we expressed the phase noise and amplitude noise in terms
of simple time-dependent functions, denoted by a(t) and p(t). In reality, to address
the nature of noise properly some statistical tools are necessary, and the concept of a
random process needs to be introduced. A few definitions are given below to establish the
vocabulary. The reader is encouraged to study the subject using appropriate references,
among which I prefer [32, 741.
10 Phase noise and frequency stabiliw 1.3 Elements of statistics 11

Basic definitions The statistical expectation E{x) is the extension of the average to stochastic processes.
It is defined as
Random or stochastic process
A mndonz process is defined through a random experiment e that associates a time- xf,(x) dx (expectation) , (1.28)
domain function x,(t) with each outcome e. The specification of such an experiment,
together with a probability law, defines a random process x(t). Each random process has where fx(x) is the probability density of the process x(t). Of course, the expectation can
an infinite number of realizations, which form an ensemble. A realization, also called be calculated for all the parameters of interest.
a sanzplefiinction, is a time-domain signal x,(t). For short, the subscript e is dropped
whenever there is no ambiguity or no need to refer to a specific outcome e. Stationarity and repeatability
A random process and its associated ensemble are powerful mathematical concepts, A random process x(t) is said to be statioiza~y(in a strict sense) if all its statistical
but they are not directly accessible to the experimentalist, who can only measure a finite +
properties are invariant under an arbitrary time shift, that is, x(t) -+x(t t'). Of course,
number of realizations. strict-sense stationarity can only be assessed a priori on the basis of physical laws. A
less general invariance property is wide-sense stationaritJ~.This definition requires that
Mean, time average, and expectation 1. the statistical expectation E{x(t)} is constant,
In the measurement of random processes (subsection 1.3.2) we use simultaneously three
2. the autocorrelation function R(tf,t") = E{x(t')x(t")) depends only on the time dif-
types of "average," the simple mean, the time average, and the mathematical expectation.
Hence, for clarity we need different notation for these.
+
ference z = t' - t" ; thus, R(z) = E{x(t z)x(t)).

Given a series of N data xi, the simple mean of x is denoted by angle brackets: Stationarity is closely related to the empirical concept of repeatability. In fact, repeating
an experiment means doing it again with the same equipment and under the same
1 conditions, yet inevitably at a different time. The 1-epeatabi1ity is expressed by the
( ~ ) ~ = - C x i (mean) interval containing the spread of measured values when the experiment is repeated a
j=i
number of times.
The simple mean is often used to average the output stream of an instrument. The
quantity x is unspecified. For example, we can average in this way a series of numbers, Cyclostationa~yprocess
a series of spectra, etc. A random process x(t) is cyclostationary of period T if all its statistical properties are
The time average of x is denoted by an overbar: +
invariant under a time shift of n T, that is, x(t) -+ x(t rz T) for integer n . The property
of cyclostationarity is an extension of the concept of periodicity and therefore is of
7-12 paramount importance in radio engineering, where the information is carried by strong
x(t) d t (time average) (1.26)
periodic signals. And of course it is of paramount importance in oscillators.

In the case of causal systems, where the response starts at t = 0, the integration limits can E~figodicityand reproducibility
be changed from -T/2 and T/2 to 0 and T. In most cases, the readout of an instrument The process x(t) is ergodic if all its statistical properties can be estimated from a single
is of the form (1.26). This means that the input quantity x is averaged uniformly over realization x,(t). In practice, ergodicity maltes sense only for stationary or cyclostation-
the time T. ary processes. It makes random processes accessible to the experimentalist. In fact, in
More generally, the time average includes a weight function w(t): the case of an ergodic process, the ensemble average is equal to the time average of a
00
single realization of the process, which can be measured. The most important averages
x (t) w(t) d t (weighted time average) (1.27) are summarized in Table 1.1.
The concept of reprodzicibility is closely connected with ergodicity. In physics, repro-
with ducing an experiment means that the same experiment is done by different operators in
different locations but using a close copy of all the relevant parts. Of course, we expect
that all the replicas of the experiment give the "same result." Using the vocabulary
of statistics, each replica of the experiment is a realization (or sample) picked from
In theoretical discussions the definition (1.27) is generally adopted as the deJinition of the ensemble. The reproducibility is expressed by the spread of values of a replicated
the nzeaszire of x. The readout of sophisticated instruments can be of this type. experiment.
12 Phase noise and frequency stabiliw I .4 The measurement of power spectra 13

Table 1.I Relevant statistical and time-domain averages. With minor changes, the integration limits can 1.3.2 The measurement of random processes
be set to 0 and T
With reference to Table 1.1, most quantities measured in a laboratory are estimates of the
Parameter Statistical expectation Time-domain average expectation E{p) of a function p associated with a random process x(t). Such measured
quantities could be the average, the variance, the autocorrelation function, etc. We rely
AVG (dc level)
on the assumption that x(t) is stationary and ergodic with regard to p . However complex
the mathematical background might be, the experimental process turns out to be simple.
We follow the steps detailed below.
second moment (power) 1. We collect N measures p i , that is, N averages of the q~~antity
p on time slots of
duration T centered at ti :

VAR (ac power)


The time slots can be either far from one another, contiguous, or partially overlapped.
For each random process, the number of degrees of freedom is related to N and to
the dead time between the measurements.
auto-correlation function R(r) E{x(t) x(t + r)} x(t)x(t +r) = 2. We use the simple mean ( P ) ~which
, is an average over the realizations of the process,
as an estimator of the quantity p :

3. We trust the mean as an unbiased estimator:


Example 1.2. Thermal noise of a resistor. Suppose that we measure the variance a 2
of the thermal voltage v(t) across a resistor R in a bandwidth B, while the resistor
is in thermal equilibrium at a temperature T. We want to establish the properties of The expected discrepancy decreases on increasing the number of degrees of freedom.
the random process associated with this experiment. In terms of statistics, the choice
of a particular resistor and of a specific time interval (tl, t2) is the outcome e of the
random experiment e. The waveform ve(t) that we record for a duration from tl to t2 is "13.3 Noise in RF and microwave circuits
a realization. The measured variance a: is the second moment of ve(t). A physical experiment can be trusted only if it is repeatable and reproducible. Therefore,
First, we restrict our attention to a single resistor and identify a sub-process, its thermal a lack of stationarity is often regarded as the result of a mistake. Non-stationary results
noise. We measure the latter a number of times for the same set of conditions. All the are found, for example, if the experiment or its boundary conditions vary in time. Radio-
measurements, which differ only in the time interval, will give the result a' = 41zTRB, frequency circuits are no exception to this rule. Yet, often the word "stationary" should
the same within a confidence interval. So we can state that the sub-process (the thermal be replaced by cyclostutionur~~. This is typical of large-signal circuits, dominated by a
noise of the specific resistor) is stationary. strong carrier. In such cases some hidden stationary process modulates the carrier, thus
Second, we measure many resistors with the same resistance R in the same conditions the visible process is cyclostationary. In some more complex cases, the noise and carrier
and simultaneously. Suppose that in all cases we obtain the value a2= 4kTRB, the same interact because the carrier modulates the noise parameters. This will be addressed in
as that obtained using one resistor on successive time slots. This means that the time Chapter 2 (on phase noise in semiconductors and amplifiers).
average is equal to the ensemble average and hence that the process is ergodic.
Third, we observe that the law o2= 4kTRB holds for all resistors at any time, thus it
describes the entire process. Thus the process is stationary and ergodic. I .4 The measurement of power spectra
Finally, we notice that the variance a2decreases from 4kTRB beyond the cutoff fre-
quency v = 1zTlh at which the photon energy hv equals the thermal energy IzT. An The most familiar example of a power spectrum occurs when white light is turned into
appropriate decrease (roll-ofl) is necessary for the total power to be finite. a continuous distribution of rainbow colors by a prism. This experiment can tell us how
the light power is distributed on the frequency axis or on the wavelength axis. Another
14 Phase noise and frequency slabiliv 1.4 The measurement of power spectra 15

J@ 1 etc.
Theoretical background
While the physical approach to the power spectrum is intuitive and straightforward, the
formal definition requires a complex theoretical framework that combines generalized
Fourier analysis and the theory of random processes. We provide only some relevant
flicker
results, and advise the reader to refer to specialized books for the details. The author's
\

favorite references are [9, 56, 751.


In the case of a deterministic signal, the spectrum is simply defined as the squared
modulus of the Fourier transform. In the case of random signals, things are a little
different. The power spectrum S(w) of a random process x(t), here called the two-sided
power spectrum and denoted by srr(w), is formally defined as the Fourier transform of
the autocorrelation function R(r):
Figure 1.4 Power spectral density S(f ). The PSD can be regarded as the power
P = l i m ~ +B-'~ J:-+$ S(f ' ) df'. The noise types (white, flicker, etc.) are those commonly
encountered in the measurement of oscillator phase noise.

example familiar to the experimentalist is the spectrum analyzer. This instrument is a


tunable filter that measures the power in a bandwidth B centered at a frequency f , which +
R(r) = ~ { x ( t ) x ( t r)} (definition of R(r)) . (1.36)
is swept.
We restrict our attention to the case of a stationary and ergodic random process x(t),
As aphysical quantity, the one-sided power spectral density7 ST(f ) associated with a
accessible through a realization x(t). All averages can be calculated from the realization
random process x(t) can be operationally defined as a distribution8 having the following
x(t); thus
properties.

1. The integral over the interval (a, b) is the signal power in that interval: R(r) = lim - (1.37)

The direct measurement of R(z) is often difficult. A more convenient way to estimate
S(w) is based on
This expression enables us to extract a quasi-monochromatic beam from white light
and gives the basis of the monochromator. Moreover, we can analyze the light,
describe the dominance of the various colors, and identify the signature of atoms,
ions, and molecules. This is the foundation of spectroscopy. which follows from the Wiener-Khintchine theorem for ergodic stationary processes.
2. The power in two separate frequency slots ( a , b) and (c, d) adds up: The function XT(w) is the Fourier transform of the truncated signal xT(t):

3. The integral over all frequencies is the total power of the signal: where the integral limits indicate a finite observation time for x(t).
According to the rules given in subsection 1.3.2 above, the power spectrum is measured
by averaging rn realizations:

This is necessary for energy conservation.


In a simplistic way, S, (f) can be regarded as the average power of x(t) in a 1 Hz
bandwidth centered at f , yet measured on an ideally narrow frequency slot in order to
remove the effect of slope (Fig. 1.4). One-sided and two-sided spectra
The Fourier transform-inverse-transform pair
If there is no ambiguity in the identification of the process, the subscript can be dropped. 00 00
In mathematics, the distribution is an extension of the concept of a function. A distribution can only be x(t)ejlLtdt + (t)= ~ ( ( u ) e j ~ ~ d w (1.41)
defined through an integral, which is our case. 2n -,
16 Phase noise and "frequency stabilifty 1.4 The measurement of power spectra 17

yields naturally a frequency representation on negative and positive frequencies. As the Table 1.2 Physical quantities associated with the spectral representation of a random voltage x(t),
actual waveforms r (t) are real signals, X(o) has the following symmetry properties: where T is the measurement time

( =X - ) (the real part is even) (1.42) Quantity Physical dimension Purpose

s{X(w)} = -3{X(-o)} (the imaginary part is odd) (1.43) Xi(f) V/Hz theoretical discussion

S'(f) = -lxT(f)l2: f >0 V 2 / ~ or


z W/Hz measurement of noise level
Ix(@)l = IX(-w)l (the modulus is even) (1.44) T (power spectral density)
arg X(w) = - arg X(-a) (the argument (phase) is odd) (1.45) 1 2
-s'(f) = ,lxi(f)12, f >0 V2 o r w power measurement of
T carriers (sinusoidal signals)
Thus, the two-sided power spectrum is an even function of w,

In some literature, as well as in the technical documentation of some instruments, one


and, as a consequence, tlze negative-fi-equency haIf-plane is redundant. A spectrum finds the following terminology:
analyzer displays the one-sided spectrum ~ ' ( w ) : 2
S'( f ) = 1 x T (f )I2 (power spectral density (PSD), v2/Hz) (1.50)

where the factor 2 is introduced to preserve the total power. However, virtualZy all the However, in most general literature the terms spectrum, power spectrum, and power
literut~n-eon tlzeor~eticalconce~~ns
uses the two-sided iflepresentation,with positive and spectral density are considered synonymous and used interclzangeably. We advise that
negative frequencies, because the formalism is clearer. the experimentalist relies on the measurement unit (v2/Hz or v2), rather than on the
The relationship between S' and S" is imposed by the equivalence of the power in terms.
the frequency interval (a, b): There are two basic types of spectrum analyzer for electrical signals, the traditional
RF or microwave spectrum analyzer and the FFT analyzer. As the technology pro-
gresses, more digital technology is included in the RF or microwave instruments and the
difference becomes smaller.

EW or microwave spectrum analyzer


This type of instrument is a superheterodyne receiver, swept in frequency, followed by
As a general rule, we prefer w for theoretical discussion, and f or v for experiments, a power meter. The sweep signal must be slow enough for the intermediate-frequency
and thus w for two-sided spectra and f or v for one-sided spectra. Accordingly, we use (IF) filter to respond stationarily. The physical quantity measured by this instrument is
the power P dissipated by the input resistance Ro in the bandwidth B centered at the
w, ~ " ( o ) in theoretical contexts , frequency v, and is expressed in watts. The bandwidth B is determined by the IF filter,
v , S' (v), f, S' (f) in experimental contexts , which can be selected from the front panel. The following cases deserve attention.

dropping the superscripts I and II unless there is a risk of ambiguity. Pure sinusoid offrqziency vo and power Po. The analyzer displays a (narrow)
lorentzian-like distribution of width B centered at the frequency vo. The peak
is at the power Po, while the actual shape of the distribution gives the frequency
1.4.3 Spectrum analyzers response of the IF filter. Thus the displayed linewidth depends on the choice of B,
while the displayed vo and Povalues are those of the input signal.
Table 1.2 shows the physical quantities associated with the spectral representation of
White noise ofpower spectral density N. The analyzer displays a power level P =
xT(t). The Fourier transform is preferred in theoretical contexts but the one-sided power
NB, which depends on the choice of B.
spectral density S' (f) = (21 T ) Ix r ( f )12 is most suitable when measuring the noise
spectrum. The one-sided quantity (21 T2)I xT (f) l2 is preferred for the power measure- In principle, it is always possible to convert the reading (the total power in the bandwidth
ment of carriers, i.e. sinusoidal signals. B of the IF filter) into the power spectral density by dividing by B. In old fully analog
18 Phase noise and frequency stahiliv 1.5 Linear and time-invariant (LTI) syste

instruments there can be a loss of accuracy due to the large uncertainty in B. Furthermore width B and height T:~,,while white noise of power spectral density N (v2/Hz)
the filter bandwidth can change with frequency and temperature, and it can age with is shown as a horizontal line at q:, = NB.
time. Conversely, in modern instruments B is defined numerically after digitizing the
signal, for there is no accuracy degradation due to the change in unit.

FPT spectrum analyzers Linear and time-invariant (LTI) systems


In this type of analyzer, the spectrum is inferred from a digitized time series of the input
signal using System function

sr(f) = ,
2
1x,(f)l2 , f >0 (1.52)
The Fourier transform is therefore replaced by a fast Fourier transform (FFT). The reader
The analysis of a physical system can be reduced to the analysis of its input-output
transfer firnetion, also called the system filnction, here denoted by the operator L:
vo(t) = L{vi(t)).The system is assumed to be (at least locally) linear, so that
should refer to [14] for a detailed account of the discrete Fourier transform and the FFT L{avi) = aL{vi) (linearity) ,
algorithm.
Hard truncation of the input signal yields the highest frequency resolution for a given L{Vl + 212) = L{v1) + L{'u2)
T and a given number of samples in the time series. The problem with hard truncation and time invariant, i.e. governed by parameters that are constant in time, so that
is that the equivalent filter has high secondary lobes, owing to the Fourier transform
of the rectangular pulses. In consequence, a point on the frequency axis is polluted by vo(t) = L {vi(t)} + vo(t + t') = L {vi(t + t')} (time invariance) . (1.54)
the signal (or noise) present in other parts of the spectrum. This phenomenon is called
Such systems, called linear and time-invariant (LTI) systems, have the amazing property
"frequency lealcage." The leakage is reduced by replacing hard truncation with a weight
that they can be described completely by their response, denoted by h(t) to the Dirac
function that in this context is called a window. The most popular such functions are
delta function S(t). Therefore, the output from an arbitrary input is
the Hanning (cosine) window, the Bartlett (triangular) window, and the Parzen window.
Hard truncation of the observation time gives a flat-top window. The choice of the most (convolution)
appropriate window function is a trade-off between accuracy and frequency lealcage that
depends on the specific spectrum.
The FFT, inherently, is linear in amplitude and frequency even if the spectrum is
displayed on a logarithmic scale. The bandwidth B is determined by the frequency span and in the frequency domain
and by the number of points of the FFT algorithm that are implemented. A typical
analyzer has 1024 points. The lower 801 points are displayed, while the upper 223 points
are discarded internally because they have been corrupted by aliasing. Thus, for example, note that an upper-case letter denotes the Fourier transform of the corresponding
using a span of 100 kHz the bandwidth is 105/800 = 125 Hz. lower-case function of time. Figure 1.5 summarizes some facts about the transfer
The measured spectrum can be displayed in two ways, listed below, without loss of function.
accuracy in the conversion. Another relevant property of LTI systems is that the exponential eJot is an eigen-
Power spectral density S' (f) = (2/ T ) Ix T (f ) l2 (v2/Hz). This choice gives a function, hence the system response to vi(t) = ejo' is vo(t) = Cejo'. This follows im-
straightforward representation of the noise. White noise of power spectral density mediately from (1.57). It turns out that the complex constant C is equal to H ( j o ) .
N (v2/Hz) is shown as a horizontal line of value N. Conversely, a pure sinusoid Hence
of amplitude &, and frequency appears as a (narrow) lorentzian-lilce dis-
tribution of width B centered at the frequency fo. The peak is equal to c:,/~,
while the actual shape of the distribution is determined by the instrument's internal The function H ( j o ) is often plotted as 20 loglo I H ( j o ) ( (dB) and arg H(jw) (degrees
algorithm. or radians) on a logarithmic frequency axis; this is referred to as a Bode plot.
xr
Power spectrum (1/ T)s'( f ) = (2/ T2)I (f) j2 (v2). The spectrum is displayed as When the input is a random process, we have to replace the Fourier transforms V,(jo)
in an RF spectrum analyzer but with the trivial difference that the power is given and Vo(jo) by the power spectra Si(w) and So(@).Thus, (1.57) turns into
in v2instead of in W. The input resistance is assumed to be 1 R unless otherwise
specified. Thus a sinusoidal signal of amplitude V,, is shown as a distribution of
Phase noise and frequency stabiiib 1.5 Linear and time-invariant (LTB) systems 21

time
s (0 h(t) becomes
LTI

1
e 0
domain system 00

X u )= ( t ) e tt for r(t) = 0, 'v't c 0 (causal signal) . (1.61)


.
vi(t)
- LTI
system .
vi(t).h(t)

Replacing jw by s = o + jw, we get the Laplace transform

Fourier
ejw t
0
LTI
system .
H(jw)e j w
X(s) =
1" x(t)e-" dt (Laplace transform) .

transform Thus, the Laplace transform differs from the Fourier transform in that the frequency is
allowed to be complex, while the strong-convergence problem arising from the lower
Vi 0 ' ~ ) Hum) Vi(jw)
0.
LTI @ integration limit t = -oo is eliminated by the assumption that x(t) = 0 for t < 0. A
system
short summary of useful properties of and formulae for the Laplace transform is given
in Appendix A.
The main reason for preferring the Laplace transform is that it gives access to the
Laplace
transform
.
e st -
LTI
system
H(s) esf
e
world of analytic functions. Analytic functions have the very important property of
being completely determined by the positions of their roots (poles and zeros) on the
-
complex plane and by the pole residues. This simplifies network analysis and synthesis
Vi($1 H(s) Vi (s)
0 LTI 0
dramatically.
system

Slow-varying systems
noise
spectra
.
Si (w) -
LTI
system .
IH ( ~ W ) Is i ( o ) The system function h(t), as any other physical quantity, can only be observed over
a finite time, which we denote by T; of course, this observation time must be long
Figure 1.5 Some relevant properties of the transfer function in LTI systems.
enough to include (almost) all the energy of h(t). Accordingly, the Fourier and Laplace
transforms can only be evaluated over the time T:

It is worth mentioning that the system function H(jw) is always a two-sided Fourier
transform even if the one-sided spectra Si and So are preferred. This is obvious from
(1.59), because the use of single-sided spectra introduces a factor 2 into both Si(w) and
So(w),and thus H(jw) must not be changed.
From this standpoint, a time-varying system can be seen as time invariant, provided
that the time scale of the fluctuation is sufficiently slow compared with the observation
time T. In the case of the Laplace transform, we can describe a slow-varying system in
Laplace transform
terms of rflootsthatflictziate or drij? on the conzplexplane. This is an appropriate way to
The Laplace transfornz is generally preferred to the Fourier transform in circuit theory, describe a fluctuating or drifting resonator.
and it is extensively used in Chapters 4 and 5. The system functions are talten to be Generalizing the finite observation time consists of introducing in the integral a
causal, that is, h(t) = 0 for t c 0. This expresses the obvious fact that the output is weight function w(t) of duration T, which sets the measurement time and has more or
zero before the input stimulus. Thus, denoting the system function by x(t), the Fourier less smooth boundaries. Thus, the Laplace transform of the weighted signal is
transform
H w ( j a )= 1 00

Jz(t)w(t)e-jut d t ,

x(t)e-j"' d t (Fourier transform) (1.60) H,~(s)=


if" h(t)w(t)e-"dt.
Phase noise and frequency stabiliu 1.6 Close-in noise spectrum 23

It may be remarked that in signal analysis (and in Section 1.4) it is often preferable Table 1.3 Most frequently encountered phase-noise processes
to set the integration time from -T/2 to T/2 in order to take full advantage of a
Law Slope Noise process Units of by
number of symmetry properties. Conversely, in the analysis of system functions the
lower integration limit is set at t = 0 because the response of the system starts at t = 0. bof 0 white phase noise rad2/Hz
b-1 f-' -1 flicker phase noise rad2
bP2f -2 -2 white frequency noise rad2 Hz
(or random walk of phase)
Close-in noise spectrum b-3 f -3 -3 flicker frequency noise sad2H Z ~
b-4 f-" -4 random walk of frequency rad2 Hz3
The spectrum of a cloclc signal is broadened by noise. Experimental observation indicates
that the power associated with the iiiost interesting noise phenomena is clustered in a " For brevity, convention dictates that the coefficients bi are all given
very narrow band around the carrier frequency. Conversely, the oscillator noise is well in r a d 2 / ~ instead
z of in their correct units. The unit rad2/H, refers to
S, (I Hz); see (1.70).
characterized by the low-frequency processes that describe the amplitude and the phase or
frequency fluctuations. This introductory section provides a short summary of classical
material available in a number of references, among which I prefer [89, 6 1, 102, 1031. The definition (1.69) does not discriminate between amplitude-modulated (AM) noise
and phase-modulated (PM) noise.
In practice, 2( f ) is always measured with a phase detector, which down-converts
Phase noise the carrier to dc. Hence, the definition (1.69) is experimeiitally incorrect because the
The most frequently used tool for describing oscillator phase noise is upper and lower sidebands overlap.
With the definition (1.69), significant discrepancies between 2( f ) and S,( f ) can
S,( f ), defined as the one-sided power spectral density of the random phase fluctua- arise in the presence of large noise.
tion ~ ( t ) .
Interestingly, the 1988 version of the IEEE standard 1139 [47] reports the obsolete
The physical dimension of S,( f ) is r a d 2 / ~ zAnother
. tool of great interest is the phase- definition (1.69), warning that it is to be replaced by (1.68). In the 1999 version of the
time power spectral density same standard [103], the obsolete definition is not even mentioned.
1

Power law
The above relationship follows immediately from the definition of the phase time,
x ( t ) = q(t)/(2n vo). The physical dimension of S,( f ) is s 2 / ~ zi.e.
, s3. A model that has been found almost indispensable in describing oscillator phase noise
Manufacturers and engineers prefer the quantity Y ( f ) (pronounced "script el") to is the power-law function
s, (f 1, where 0

1
Y(f)=ZS,(f) (definitionofY(f))
S,(f) = C bif
i=-4
-
(or less)

is given in dBc/Hz, i.e. the quantity used in practice is Phase-noise spectra are (almost) always plotted on a log-log scale, where a term f'
maps into a straight line of slope i, i.e. the slope is i x 10 dB/decade. The main noise
processes and their power-law characterization are listed in Table 1.3. Additionally,
and is given in dBc/Hz. The unit dBc/Hz stands for "dB below the carrier in a 1 Hz Table 1.4 provides a number of useful relationships discussed in Section 1.7.
bandwidth." The IEEE standard 1139 [103] recommends reporting the phase-noise In oscillators we find all the terms in (1.70) and sometimes additional terms with
spectra as Y (f ). Historically, Y (f ) derives from early attempts to measure the oscillator higher slope, while in two-port components white noise and flicker noise are the main
noise with a spectrum analyzer, for it was originally defined as follows: processes. In fact, the phase noise of a two-port device cannot be steeper than l/f
one-sideband noise power in 1 Hz bandwidth for f -+ 0, otherwise the group delay would diverge rapidly. This is why higher-slope
Y ( f 1 =: (obsolete definition) . phenomena in amplifiers are in fact "bumps" that appear superposed on the fliclter noise.
carrier power
(1.69) For example, this is the case for the l / f 5 noise of thermal origin, to be discussed in
The definition (1.69) has been abandoned in favor of (1.68) for the following reasons. Chapter 2.
1.6.3 Frequency noise 1.7 "Tinne-domain variances
Another way to describe oscillator noise is to ascribe its randomness to the frequency
fluctuation (Aw)(t) or (Av)(t) instead of to the phase (see (1.15) and (1.16)). The In this section we use the three types of average introduced in Section 1.3, namely: the
time-domain derivative maps into multiplication by j w in the Fourier transform, thus time-domain average, denoted by an overbar on the variable, as in 7; the average of N
into multiplication by w2 in the spectrum. Hence, temporarily using for the Fourier values, denoted by ( ) N ; and the mathematical expectation, denoted by E { }. Some care
frequency and w for the carrier, it holds that s A m ( a )= Q~s,(Q) and thus is necessary to avoid confusion.
The reading VIc(r)of a frequency counter9 that measures the frequency v(t) on a time
S ~ v ( f=
) f2Sq(f), interval of duration t starting at time k t represents the time average of v(t):
because Av = (1/2n)Aw, and SAv= (1/4n2)sAU.
The use of SAv(f ) is common in the context of lasers, while SJ,(f ) is more com-
mon in radio-frequency metrology. Recalling the definition of the fractional frequency using the normalization TIC= (TIC- vo)/vo, giving
This is converted into h(t)
fluctuation y(t) = (Av)(t)/vo, Eq. (1.7 1) becomes

In practice, the resolution of a frequency counter is often insufficient. When this is


Of course, the power-law model also applies to SA,(f ) and to S,,(f ) . In the literature the case, it is convenient to measure a low-frequency beat note vb obtained by down-
the coefficients of SJ,(f ) are denoted by hi. Comparing (1.70) and (1.72), we find converting the main frequency vo with an appropriate reference. The resolution improves
2 by a factor vO/vb.This experimental trick does not affect the development below.
sj,(f) = C hif'
i=-2
Given N samples Yk(t), the classical variance is

with

In the presence of fliclcer, random walk, and other phenomena diverging at low frequen-
cies, the classical variance VAR depends on N and on t , which remain to be specified.
In a comparable way to that for the power-law model of S,( f ), it is generally accepted It also has a bias that grows with large measurement times. In the quest for a less biased
that the coefficients hi are each given in HZ-', corresponding to the physical dimension estimator, the Allan variance and the modified Allan variance have been developed.
of S,,(f), instead of the unit appropriate to each one.
1.7.1 Allan variance (AVAR)
Amplitude noise The most often used tool for the time-domain characterization of oscillators is the Allan
variance (or the Allan deviation). The Allan variance
The effect of amplitude noise in oscillators may be not negligible, for a variety of
reasons. An example is the sensitivity of the resonant frequency to the power in the
piezoelectric quartz, due to the inherent nonlinearity of the quartz lattice and laown
as the "isochronism defect" [42]. Another example is the effect of radiation pressure is defined as the expectation of the two-sample variance, i.e. the classical variance (1.77)
in cryogenic sapphire resonators [18], which turns the amplitude noise into frequency evaluated for N = 2. The deviation is the square root of the variance:
noise.
A power law is suitable for describing the amplitude noise of the clock signal and that
of a two-port component as well. In all cases, the dominant phenomena are the white It is assumed that the two samples are contiguous in time. If the samples jilc(r) are not
(f ') noise and the fliclcer (11f ) noise. Bumps can be present in some portions of the contiguous then a correction is necessary, which depends on the noise type [3].
frequency axis, but the asymptotic law cannot be steeper than l/f at low frequencies
otherwise the signal amplitude would diverge rapidly.
The reader should be warned that some frequency counters use a triangular weight function [80,25], instead
A detailed discussion about AM noise is available in reference [79]. of the uniform weight function needed here.
26 Phase noise and frequency stability 1.7 Time-domain variances 27

AVAR

Figure 1.6 Transfer functions I H A ( jf ) l2 and I HM(jf )I2 of the Allan variances.

In practice, the statistical expectation is replaced by the simple mean. Given a stream of
MVAR

= time

Figure 1.7 Weight function of the Allan variances u j ( r ) and mod o j ( r ) . Reproduced from [80]
M contiguous samples &(t), we have M - 1 differences - &.Thus the measured with the permission of the AIP, 2007.
Allan variance is
where the weight function (Fig. 1.7) is

Frequency-domain interpretation t<t<22,


The Allan variance (1.78) is equivalent to a filter or transfer function H ~ ( j f ) ,see
elsewhere .
Fig. 1.6, applied to S,,(f):
This is similar to a wavelet variance; however, the weight function wA(t) differs from
that for a Haar wavelet in that it is normalized for power instead of energy. In fact, the
energy1' of wA,

is proportional to 1I t , while the energy of a wavelet is equal to 1, by definition.


This filter has an equivalent noise bandwidth of half an octave, having a main lobe with Interestingly, in (1.84) o i ( r ) has the structure of a squared scalar product. The weight
a peal<at f t 0.3 7. Consequently,
function wA(t) is odd with respect to t = t (Fig. 1.7), and thus the Allan variance
AVAR detects only the odd part of the input signal. However, the noise power is equally
divided into odd functions and even functions. Consequently, AVAR detects half the
noise power. This is related to the fact that E{wA}= l / t , while the response to the
white noise S,,(f) = ho is o$(r) = ho/(2r).
thus the filter response to a white noise S,,(f ) = ho is o$(r) = hol(2r).

Wavelet interpretation Modified Allan variance (MVAR)


Writing &(r) as the integral (1.76), the Allan variance can be given as The modified Allan variance MVAR [96] (see Fig. 1.7) was introduced to improve the
"behavior" of AVAR in the presence of fast noise processes. It is defined, for t = nto, n

lo Applying statistics to physics, the concepts of power and energy inevitably relate to physical quantities,
expressed in joules or watts or dimensionless mathematical entities, as in (1 36).
28 Phase noise and frequency slablliiw 1.8 Relationship bebeen spectra and variances 29

an integer, as is proportional to l / t . Interestingly, wM(t)is an odd function with respect to t = 3212,


and thus MVAR detects only half the noise power. The same fact was observed in the
case of AVAR.

(1.87)
AVAR versus MVAW
The deviation is the square root of the variance:
The modified Allan variance improves on the Allan variance in that it discriminates
mod o;,( r ) = d m o d o: ( r ) =
R- (MDEV) . (1.88) between white PM noise and fliclter PM noise. However, for a given measurement time
Once again, the statistical expectation is replaced by the mean of contiguous values, t the measurement of inodo,,(r) taltes longer than o;,(r). The reason is that the weight
trusting the ergodicity of the process. function in (1.78) taltes a time 22 and the weight function in (1.87) taltes a time 3 t .
A problem with both variances is that, for the same noise process, o i ( r ) differs from
Frequency-domain interpretation mod a;?(r). In general mod o j ( r ) < o;(r); the actual relationship depends 011 the noise
In the frequency domain, the MVAR computation is equivalent to the application of a process. This goes with two facts,
filter or transfer function H M (fj ) applied to S,,(f):

and
whose shape depends on n. On increasing n, the shape converges rapidly to
sin6 itrf
lim
--+00
1HM(jf)i2 = 2 (Fig. 1.6) .
( n r f l4
11
nto=t

For n 2 10, the asymptotic form is accurate enough for most practical purposes. This Thus, the AVAR response to a white frequency noise S,,( f ) = ho is o i ( r ) = 1z0/(2r),
filter has an equivalent noise bandwidth of a quarter octave, with a peak at f r % 0.3 1 while the response of MVAR to the same noise is mod o i ( t ) = ho/(4r).
[131.

Wavelet interpretation Relationship between spectra and variances


The modified Allan variance can also be interpreted as a wavelet variance
Figure 1.8 and Table 1.4 provide a summary of the relationship between the spectra
and the Allan variance. The relationship between S,( f ) and S,,(f ) is exact, thus it is
always possible to convert between S,( f ) and S,,(f ) in both directions without loss of
in which the weight function converges to information.
In contrast, the conversion from spectra to variances is always approximate because
we cannot evaluate the integrals (1.8 1) and (1.89) completely. Moreover, the conversion
from o,'(r) to S,,( f ) is not free from errors in the general case [39]. This is related to
the presence of side lobes at f t 1.5,2.5, . . . in the band-pass filter 1 H A ( fj ) l2 (see
1 (1.82) and Fig. 1.6), which mix up the spectrum. A similar phenomenon occurs in the
(t - 3 r ) 2r c t < 3 r , modified Allan variance (1 27).
A further problem is that the variances detect only the components of the signal that
elsewhere
have odd symmetry, with respect to t = r in the case of AVAR and to t = 3r/2 in
for ro << r , or equivalently for n >> 1 (see Fig. 1.7). Once again, the weight function the case of MVAR. This blindness to even functions of time maltes the variance phase
differs from a wavelet in that it is normalized for power instead of energy. In fact, the sensitive when the signal is a deterministic modulation, such as the 50-60 Hz mains
energy of the weight function, supply. Of course, spectral analysis has no preference for odd or even signals.
That said, it is clearly unfortunate that o,,(r) has been adopted as the standard for the
characterization of clocks rather than S,,(f ). However, q,,(r) and mod o,,(t) incorporate
Phase noise and frequency stabiliw L.9 Experimental techniques 31

\ :, random walk freq. Table 1.4 Noise types, power spectral densities, and Allan variance

Noise type s,(f> s~(f> sves~ .)?(r)" mod o;(r)"


', flicker fseq.
;b-, f -3 white PM

b-1
flicker PM b-lf-' hlf h i = -F +
[1.038 3 1n(2nfHt)]
vo hl r-2
X --------
(2n12
b-2 1
white FM b-2 f - 2 ho ho = -, - ho z-'
vo 2

linear frequency
drift y
-

" fH is the low-pass cutoff frequency, needed for the noise power to remain finite.

3. The dynamic range is insufficient for the instrument to detect the noise sidebands; it
is "dazzled" by the strong carrier.
4. The noise and the frequency fluctuations of the local oscillator are larger than the
noise of most good oscillators of interest.

The first two problems are solved by smart digital technology in some modern analyzers,
which are capable of phase-noise measurement. However, the dynamic range is still
Figure 1.8 Power laws for spectra and Allan variance. limited by the available technology of frequency conversion. The higher dynamic range
requires a carrier cancellation method, such as direct phase measurement with a double
balanced mixer or a bridge, which cannot be implementedll by a spectrum analyzer.
the effects of spurs, which are difficult to evaluate from the spectra. Finally, o,(r) was Finally, the spectrum analyzer scans the input range with a voltage-controlled oscillator.
easier to measure with the early instruments. Owing to the wide tuning range, the spectral purity of such an oscillator is insufficient
to measure the phase noise of most fixed-frequency oscillators. In conclusion, even if
some modern spectrum analyzers are capable of measuring phase noise, this function is
generally inadequate in nunierous practical cases.
Experimental techniques

Spectrum analyzer
The scheme on which virtually all commercial phase-noise measurement systems are
When used to measure a quasi-perfect sinusoid like the clock signal (l.6), the traditional
based includes a phase-loclted loop (PLL), as shown in Fig. 1.9. The double-balanced
spectrum analyzer suffers from the following limitations.
mixer is saturated at both inputs by two signals in quadrature, and it convertsthe phase
1. The IF bandwidth is too large to resolve the noise phenomena.
2. The instrument is unable to discriminate between AM noise and PM noise. Actually, a carrier suppression technique was proposed as an extension of a spectrum analyzer to measure
2( f ) [50].After the prototype, this solution seems not to have been used further.
Phase noise and frequency slabiliw Exercises 33

quadrature relationship at the double-balanced mixer ports. With such loose PLLs,
(1.98) is approximated by Su(f)/S,( f ) = lzi, from which we estimate S,( f ) .
A more sophisticated approach consists of setting up a tight phase loclting; of course,

m
the high-pass function (1.98) must be measured accurately and subtracted from the
measured spectral density. This approach has the following advantages.
analyzer
REF 1. In practice, the two oscillators are more or less loosely coupled by lealtage. This gives
the wrong impression of a phase noise lower in magnitude than the actual value. Of
course, the lealtage is hardly reproducible because it depends on the experimental
2nk& 1 phase-lock feedback layout. With an appropriate choice of the operating parameters, however, a tight loop
can override the effect of lealtage.
2. The phase noise rises as 1/f 3 , or as 1/f at low frequencies. After tight locking,
S,( f ) rises as 1/f , or as 1/f at low frequencies. This reduces the burden of dynamic
range at the FFT input.
Figure 1.9 Practical scheme for the measurement of oscillator phase noise. The most appropriate time constant is determined on the basis of experimental criteria.
A good starting point is the DUT cutoff frequency, where the l / f or l / f phase noise
difference Pi - (o, into a voltage v d = lzd((oi- yo), fed back to the VCO input and turns into 1/f 3.
detected by the FFT analyzer. In this configuration, the PLL is a high-pass filter for The assumption that the reference oscillator is ideally stable is realistic only in some
the oscillator noise. In fact, the noise measurement relies on the fact that the loclted routine measurements where a low-noise reference oscillator is available. Conversely,
oscillator does not track the fast fluctuations of the oscillator under test, so that the in the measurement of ultra-stable oscillators the PLL is used to compare two equal
error signal is available for the noise measurement. The PLL ensures the quadrature oscillators. In these conditions, the DUT noise is half the raw value of S,( f ).
relationship for the duration of the measurement. The reference oscillator is pulled to The background noise is determined by the following facts.
the DUT fiequency vo by the addition of a dc bias to the control voltage.
The white noise floor is chiefly due to the dc low-noise amplifier. The reason is the
Using Laplace transforms, denoted by the upper case letter corresponding to the
poor phase-to-voltage conversion gain of the mixer, which is about 0.1 to 0.5 Vlrad.
time-domain signal, the PLL transfer function is
White noise values of - 160 to - 170 dB rad2/Hz (bo = 10-l6 to 10-l7 rad2/HZ) are
often seen in practice.
0 The flicker noise is due to the mixer. Common values are - 120 dB rad2/Hz at 1 Hz
offset (b- = 10-l2 rad2/Hz) for microwave mixers and - 140 dB rad2/Hz (b- =
where 10-l4 rad2/Hz) for RF units up to about 1 GHz.
In some cases, the amplitude noise pollutes the phase-noise measurement through
lzL = 1 ~ ~ 1 ~
(loop
~ gain)
1 ~ ~ (1.97) the mixer power-to-dc-offset conversion [8 11. This is due to symmetry defects in real
mixers.
is the loop gain. The VCO gain lc, is given in radl(Vs). The mixer phase-to-voltage
gain kcd (Vlrad) also accounts for the gain of the dc amplifier, and kF is the gain of the
feedback-path amplifier. In order to rewrite (1.96) for the noise spectra, we take the
square modulus after substituting s = j 2 n f and obtain
Exercises

1.1 A cosine signal has frequency 10 MHz and rms amplitude 2 V. Write the clock
signal in the form (1.6) and sltetch the associated phasor.
Naively, one would be inclined to set a low loop gain kL, so that the cutoff frequency
f, = k ~ l ( 2 n )is low enough to leave the device under test (DUT) free to fluctuate at 1.2 Sltetch the complex envelope (phasor) associated with the signal of the previous
all the frequencies in the spectral analysis. Of course, kL must still be high enough exercise, after adding a phase noise of 2 x rad rms. Repeat for an amplitude noise
for the DUT to track the reference oscillator in the long term, which guarantees the of 5 x but no phase noise.
Phase noise and frequency sttabillw

1.3 The 10 MHz signal of the previous exercise, as affected by a phase noise equal to
2 x 1o - ~rad rms, is multiplied to 100 MHz and to 200 MHz using noise-free electronic
circuits. Sketch the complex envelopes (phasors) associated with the 100 MHz and 200
Phase no
MHz outputs.
1.4 Your wristwatch lags 2 seconds per day. Plot x(t) and y(t).
1.5 As in the previous exercise, your wrist watch lags 2 seconds per day. Plot p(t) and
(Av)(t) assuming that the internal oscillator is a 215 Hz quartz.
1.6 Repeat the previous exercise replacing the quartz by a swinging wheel oscillating
at 5 Hz.
Understanding amplifier phase noise is vital to the comprehension of oscillators. There
1.7 Your wrist watch is free from noise and errors. However, reading the time on it are two basic classes of noise, additive and parametric, as illustrated in Fig. 2.1. However,
takes in a random error of 0.5 s max. Plot the Allan deviation oJ,(r).Assume that the their behavior, and the underlying physical mechanisms, are strikingly different. This
error distribution is uniform in (-500, +500) ms, and that the spectrum is white. chapter focuses on amplifiers but includes a few words on the photodetector, which
1.8 A near-dc amplifier has gain 20 dB, white noise 4 nv/&, and flicker noise forms part of the resonator-amplifier loop in the emerging photonic (or optoelectronic)
4 n ~ / - at 1 Hz. Plot the noise spectrum of the input and output voltages. oscillators. The same principles apply to other electronic and optoelectronic components.
Random noise is normally present in electronic circuits. When such noise can be
1.9 Plot the Allan deviations of the input and output voltages for the near-dc amplifier represented as a voltage or current source that can be added to the signal, it is called
in the previous exercise. Tip: use Table 1.4, with v instead of y. Of course, the physical additive noise. Another mechanism can be present, referred to as parametric noise, in
unit of such a o, is V. which a near-dc process modulates the carrier in amplitude, in phase, or in a combination
1.10 Write the transfer function H(s) of a simple RC low-pass filter (R in series, C to thereof. The most relevant difference is that additive noise is always present, while
ground) with R = 1 1cQ and C = 2.2 nF, and sketch the Bode plot. parametric noise requires nonlinearity and the presence of a carrier. Additive noise is
generally white, while parametric noise can have any shape around vo, depending on
1.11 The input of the above low-pass filter is a 200 n ~ / & white noise source, the near-dc phenomenon. Additive noise originates in the region around the carrier
uniform from dc to 1 GHz. Sketch the output noise spectrum and calculate the total frequency vo while parametric noise is brought there from near-dc by the carrier. This
output power. Ignore the thermal noise added by the filter. is made evident by a gedankenexperirnent (thought experiment) in which we inspect the
1.I2 Thermal noise is governed by the law P = kT B, where lzT is the thermal energy amplifier output around vo with an ideal spectrum analyzer, switching the input carrier
at temperature T and B is the bandwidth. Such a noise is added to an ideal 1 mW on and off. A noise process such as l / f noise cannot be present around vo when the
carrier. Calculate 2?( f ) and S,( f ) at the temperatures To = 290 K, TiiquidN = 77 I<, and carrier is switched off, while the additive noise will still be there unchanged.
& p i d He = 4.2 I<. Among parametric noises, flicker (1/f noise) is so important that the term "parametric
noise" is sometimes used synonymously with it. Parametric noise originating from near-
dc white noise is generally not relevant in practice. Environmental fluctuations often
show up as parametric noise at very low offset frequencies.

Fundamental noise phenomena

Thermal noise
The power spectral density of blaclcbody radiation is described by the law
2.2 Noise temperature and noise figure 37

Example 2.1. The available noise power of a resistor at the temperature T = 290 K is
1zT = 4 x lop2' W/Hz. Assuming a resistance R = 1 kQ, the spectrum of the available
noise voltage is kTR = 4 x lo-'' V ~ / H Zthus
; d m = 2 n ~ / & . Leaving the
resistor terminals open, the spectrum is 4kTR = 1.6 x 10-l7 v2/Hz; thus d m=
4 nv/qTG.

rl
microscopic
'-4
environmental

thermal noise,
shot noise,
mi 1 (50-60 Hz)
ed.
2.1.2 Shot noise
In most electronic devices the conduction of electricity takes place as a stream of separate
electrical charges q = 1.602 x 10-l9 C (the charge on an electron), each one giving rise
etc.
to a current pulse. This is a random current i ( t ) that originates the white noise of the
Figure 2.1 Noise types in amplifiers. The same classification is valid for both phase and spectrum,
amplitude noise.
si(~)=2qi (A~/HZ), (2.6)

Energy equipartition in classical thermodynamics states that the energy is kT per degree where 7 is the average current. Equation (2.6) is independent of whether the carriers are
of freedom. Electromagnetic radiation has two degrees of freedom, electric field and the electrons or holes or a combination of these. The power spectrum of the random current
magnetic field, and so the energy of blackbody radiation is kT. Blackbody radiation is i (t) flowing into a resistor R is
white noise at low frequencies, decreasing as l/(e"v'"T - I), below the cutoff frequency Si (v) = 2qT R (W/Hz) . (2.7)
kT/ h. At the cutoff, the photon energy h v equals the thermal energy 12T. For reference,
the cutoff occurs at 6.04 THz (49.6 pm wavelength) at 290 I<, and at 87.5 GHz (3.42 mm
wavelength) at liquid He temperature (4.2 I<). The physical dimension of the thermal Flicker noise
noise spectrum is that of energy, expressed in joules. Yet in spectral analysis we prefer The existence of excess noise, later called corztact rzoise orflicker noise, was initially
the equivalent unit W/Hz. recognized by Johnson [57]; and after that it was studied in thin films [4, 51 and carbon
Resistor noise, observed and analyzed theoretically in 1928 [58, 721, comes from the microphones [19]. Later, it was observed that noise with a spectrum scaling of the type
coupling of the blackbody radiation of the resistor to the rest of the electric circuit. The 1/f " , with x E (0.8, 1.2), is present in an wide variety of physical phenomena, such as
power spectrum density is often denoted by the symbol N, taken to be independent of stellar emissions, lake turbulences, Nile flooding, and of course virtually all electronic
v, instead of S(v): devices. In the absence of a unifying theory that explains why flicker is so ubiquitous, we
rely on models, the most accredited of which are due to Hooge [49] and to McWhorter
N = 122" (resistor noise, WIHz) . (2.3)
~691.
The spectrum of the random voltage available1 at the resistor ends in V ~ / H Z
is

S, (v) = lcT R (resistor terminated) (2.4) Noise temperature and noise figure
if the resistor is impedance-matched to a load, and
The noise temperature and the noise figure, most often used in the context of amplifiers
and radio receivers, are parameters that describe the white noise of a device (Fig. 2.2).
S,(v)= 4kT R (resistor open) (2.5)
In analogy with the thermal noise kT, the amplifier noise spectrum Na is referred to
if the resistor terminals are left open. the amplifier input and given in terms of its eqt~ivalentrzoise te7~zpeiatt~r.e
Ta:
Na = lzz (noise temperature Ta) . (2.8)
In the jargon of electric networks, the term "available power" refers to the maximum power delivered by
the generator, which occurs when the generator is impedance-matched to the load. By extension, the term The amplifier is always input-terminated to some dissipative load, which contributes its
"available" also applies to voltage and to current, referring to the condition of impedance matching. thermal noise, N, = kT,. The total equivalent noise N, referred to the input is therefore
Phase noise in semiconductors and amplifiers 2.2 Noise temperature and noise figure 39

Ro Ts Tt? Ta

.
out
T, = T,+ T, (equivalent
nolse temperature)
out
0

power loss t2 power gain

Figure 2.3 Amplifier noise in the general case of non-uniform temperature.

The practical use of the noise figure F is that the resistor thermal noise Ns = kT can
out be replaced by the equivalent noise
F=-Q-- (noise figure)
+

To
Figure 2.2 Equivalent noise temperature and noise figure in amplifiers. The typical noise figure for low-noise amplifiers is 0.5-2 dB, depending on technology,
on frequency, and on amplifier bandwidth. It is higher in low-cost amplifiers and in
given by power amplifiers.
It is to be pointed out that the noise temperatzire, and the rzoisefigtire as well, account
for all noise phenomena including shot noise. For example, at high optical power the
equivalent noise temperature at the output of a photodetector is mainly due to shot noise,
In radio and microwave engineering, the noise parameter most commonly used by which has very little to do with the device's physical temperature or the temperature of
manufacturers is the noisefigure F , defined as the optical source.
Some authors describe amplitude noise and phase noise using a non-white spectrum,
F = (SNR)input (definition of the noise figure F ) , by introducing a frequency dependence into the noise temperature and the noise figure:
(SNR)o,tput thus, T(f ) and F (f). This may be wrong in the general case, where the amounts of
where SNR is the signal-to-noise ratio. In practice, the definition (2.11) is often replaced amplitude noise and phase noise are not the same.
by the ratio of Ne and the thermal spectrum associated with a reference temperature To:
Ne jt The general case of non-uniform temperature
F=- (practical noise figure F ) .
k To
Let us consider the general case of Fig. 2.3, in which the amplifier input is terminated
For the definition of F to be unambiguous, and for (2.12) to be equivalent to (2.1 l), it to a resistor through an attenuator of power loss l 2and the whole circuit is impedance-
is necessary that the amplifier is appropriately terminated to its input impedance Ro and matched to Ro. As the temperature is left arbitrary, we need to use the noise power
that the input termination is at a standard temperature To. In practice, it is preferable for spectral density instead of the noise figure. The noise of the termination, referred to the
the entire system to be at the temperature To. The standard value generally adopted is amplifier input, is

with associated power spectrum kTo = 4 x W/Hz, that is, - 174 dB m in a 1 Hz


bandwidth. because the attenuator attenuates the noise of the termination, as it does with any signal.
Substituting (2.12) into (2. lo), with T, = To, we find that The attenuator noise referred to the amplifier input is

hence From this relationship it follows that the amplifier receives a noise ltTo from the resistance
Ro it sees at its input when I;. = Te = G.
The amplifier noise contribution is Na = ( F - l)kTO,as given by (2.10). The total
and + +
noise referred to the amplifier input is Ne = N, Nl Na, i.e.
40 Phase noise in semiconductors and amplifiers 2.2 Noise temperature and noise figure 41

as Rbbl (also Tflbb' in some textboolts) in the pioneering hybrid-n model proposed by
Giacoletto [37] and later in the popular Gummel-Poon model [44]. For the salte of
completeness, the other basic noise phenomena are shot noise in the base region and in
the collector and partition noise.

kT~ (Fl - 1)kTo (F2-l)kTo (F3-l)kT0


Field-effect transistors
Figure 2.4 Noise temperature and noise figure in amplifiers. The most relevant source of white noise in these devices is the channel noise fed back
to the gate through the gate-channel capacitance and amplified 1100, Section 9.31. This
The scheme of Fig. 2.3 can be generalized further by introducing impedance mis- mechanism also accounts for the increased noise figure beyond a cutoff frequency.
matching. The analysis is omitted here. Mosfets have a low noise temperature, generally of the order of 10 I<, and capacitive
A non-uniform temperature is found, for example, in cryogenic oscillators and in oven- input impedance. A complex networlt is needed to match the gate to the standard resistive
controlled quartz crystal oscillators (OCXOs). In the first case, the resonator operates at input (50 a ) . This networlt introduces loss and thus increases the noise figure. Needless
low temperatures (4-80 I<), where it exhibits a high Q factor. In the case of an OCXO, to say, a large-bandwidth design results in a high input loss, leading to a high noise
the resonator is kept at the turning point of the frequency-temperature characteristics, figure.
which occurs at some 75-80 "C. This choice maltes the temperature control simple and
effective. In both cases, the sustaining amplifier is at a convenient temperature. High-speed solid-state photodetector
This type of photodetector is a reverse-biased junction in which the light originates a
photocurrent. A detected photon creates an electron-hole pair, +q and -q; these are
2.2.2 The noise figure of cascaded amplifiers accelerated in opposite directions by the electric field. Thanlts to the Ramo theorem, the
In a cascade of amplifiers, it is convenient to refer the total noise to the input of the first shape of the current pulses at the detector ends is such that the total charge transferred is
amplifier. With reference to Fig. 2.4, adding the individual contributions we divide the q . Given that the optical power is P(t) at frequency v, the number of photons per second
noise (F2 - l)kTo of the second amplifier by the power gain A: of the first amplifier, the must be P/(h v). The photocurrent is
noise (F3 - l)lcTo of the third amplifier by the power gain A : A ~of the two preceding
amplifiers, etc. The final result is

where q is the quantum efficiency, i.e. the probability that the photon is detected. The
quantity
Generalizing, we add the noise (F,,, - 1)Jcfi of the 171thamplifier after dividing it by the
power gain n:.':yl
A: of the preceding m - 1 amplifiers. Using the noise figure instead
417
p=-
hv
(responsivity, N W ) (2.24)
of the noise power, (2.2 1) becomes
is called the responsivity. For a typical quantum efficiency q = 0.8, the responsivity
p is about 1 AIW at h = 1.55 pm and about 0.85 N W at h = 1.32 pin. Combining
(2.6) and (2.23), the power spectral density of the detector shot noise is
which is lulown as the Friis formula [33]. The important conclusion that can be drawn
from (2.22) is that there is a strong dominance, in general, of the noise introduced by
the first amplifier over the other noise contributions. Therefore, special care should be
talten in the choice of the first stage of amplification.

Example2.2. AbeamofpowerP = 1mWatwavelengthh = 1.55 pmcarries7.8 x 1015


2.2.3 Noise in amplifiers and photodetectors photons per second. Assuming a quantum efficiency 7 = 0.8, the photocurrent i =
1 rnA. The shot noise Si= 3.2 x z thus f i = 17.9 p ~ / & . When the
A 2 / ~and
Bipolar transistors photocurrent flows into a 50 resistor, the noise is 1.6 x W/Hz or 8.01 x
In modern transistors, it turns out that the thermal noise of the distributed base resistance v2/HZ (corresponding to 0.895 n ~ / & ) .
is the most relevant noise type [loo, Section 9.41. This resistance is usually referred to
42 Phase noise in semiconductors and amplifiers 2.3 Phase noise and amplitude noise 43

No carrier Parametric
4 S(f > + S(f up-conversion

LSB USB

peak phase Figure 2.6 Parametric up-conversion of near-dc flicker in amplifiers. Reproduced from (871 and
used with the permission of the IEEE, 2007.

the two phase-noise sidebands is a vector of peak phase

Figure 2.5 The contribution of additive noise to the phase noise.

oscillating sinusoidally around the real axis. The rms phase is


2.3 Phase noise and amplitude noise
2.3.1 White noise
All noise processes with white-spectrum statistics are grouped under the label of additive The power spectral density is the average square phase noise per unit bandwidth and is
noise, i.e. a random white process added to the useful signal. Of course, the noise is independent of frequency (S,( f ) = bo), because this is white noise. Hence
not correlated with the signal, nor connected with it by some conformal transformation.
Under these conditions, the noise power is equally partitioned into the two degrees of NB
bo = - (white phase noise) , (2.29)
freedom, that is, the in-phase and quadrature components with respect to the carrier or, Po
equivalently, the amplitude noise and phase noise. In the case of thermal noise, we find
which in the case of the phase noise of an amplifier can be rewritten as
an energy equal to fk~ for each degree of freedom.
We will derive the expression for the phase noise from the noise power spectral density FlcTo
N using the phase-vector representation of a sinusoid (Fig. 2.5). The carrier power Po bo = ----- (amplifier) .
Po
corresponds to a vector of rms voltage on the real axis. The noise power in the
bandwidth B is N B and thus f N B in each degree of freedom, i.e. amplitude and phase.
Accordingly, the phase-noise upper sideband (USB) is a random sinusoid of frequency
Flicker noise
+
vo f and of rms voltage
Understanding flicker noise starts from the simple observation that the output spectrum is
of the white type - flat in a wide frequency range - when the carrier power is zero and that
the close-in noise shows up only when the carrier power present at the amplifier output
This is represented as a vector of rms voltage V,, rotating at frequency f around the tip is sufficiently large (Fig. 2.6). At closer sight, it is observed that the noise sidebands
of the carrier vector, in quadrature with the carrier at time t = 0. Similarly, the phase- grow in proportion to the carrier. The obvious consequence is that the close-in fliclcering
noise lower sideband (LSB) is a vector of the same rms voltage, rotating at frequency results from the near-dc flicker noise that modulates the carrier in amplitude and phase.
-f , and also in quadrature with the carrier at the time t = 0. The sum of the carrier and Two simple models are presented below.
44 Phase noise in semiconductors and amplifiers 2.3 Phase noise and amplitude noise 45

Nonlinear mechanism Quasi-linear parametric mechanism


The first mechanism of noise up-conversion relies upon the amplifier's non-linearity, Another mechanism of near-dc noise up-conversion is the fluctuation in the amplifier
which for our purposes can be represented as a polynomial truncated at the second gain. This is modeled by replacing the constant voltage gain A by
degree:
A(t) = A. [ 1 + n (t)] (amplifier gain) (2.37)
= Ao [l + nf(t) + jnU(t)] , (2.38)
The coefficient a1 is the voltage gain denoted by A elsewhere in this book. Given a
carrier C.;ejWot
at the input, the amplifier's noisy input signal is
+
where n(t) = n'(t) jn"(t) is the near-dc noise. It should be noted that the noise
n(t) introduced here is not the same physical quantity n(t) as that in (2.32), (2.33).
For example, in the case of a bipolar transistor, rzf'(t) is easily identified as a signal
proportional to the fluctuating voltage that modulates the collector-base reverse voltage
and in turn the capacitance. The latter affects the amplifier phase lag, which generates
where n(t) = nf(t) + jnf'(t) is the internally generated near-dc noise. The noise n(t) phase noise. Nonetheless, we will keep the analysis at a higher level of abstraction,
is the amplifier's internal noise referred to the input and accounts for the efficiency of leaving the nature of n(t) unspecified.
the modulation process. One could object that the use of the analytic signal ejwo'in 14 the presence of a carrier V,ejwotat the input, the amplifier output is
nonlinear circuits is not allowed in the general case because it describes only the positive
frequencies; the consequence is that the frequency-difference terms do not appear in the
v,(t) = A. [1 + nf(t)+ jnM(t)]KejLMt,
equations. Nonetheless, the use of ejwO' is correct in this case because we select only the from which it follows that
up-conversion term, which brings n(t) close to the carrier. Thus, combining (2.3 1) and a(t) = rz'(t) , (2.40)
(2.32) and discarding the terms far from vo, we get the output signal:
p(t) = nU(t). (2.41)
Once again, a(t) and p(t) are independent of the input amplitude V, and thus S,( f ) and
In this representation, the term V,alejWotis the output carrier and the term V,[2a2nr(t) + S,( f ) are expected to be independent of the carrier power.
j2a2n"(t)]ejmt is the close-in noise associated with the output signal. The latter orig- It is to be pointed out that the parametric model is near.& linear.. In mathematics, a
inates from the cross-product terms coming from the expansion of azv;(t) in (2.31). function f (x) is linear if it exhibits the following two properties:
It follows that, using the signal representation (1.6), the random amplitude and phase
fluctuations are f (ax) = a f ( 4 ?
(2.42)
f ( x +Y> = f ( x >+ f01) (2.43)
One can verify the property (2.42) by replacing KejWotby a v , ~ J ~ ino ~(2.39) and the
+
property (2.43) by replacing ~ , e j @by~ 'v1ejw1' V2ejw2'in (2.39). In both cases the
deviation from linearity consists of terms containing nf(t) and rz"(t), which are pro-
Interestingly, a(t) and p(t) are independent of the input amplitude 6.Therefore the portionally small in low-noise conditions, i.e. for Inf(t)l << 1 and Irzlf(t)l << 1. This
power spectral densities S,( f ) and S,( f ) are expected to be independent of the carrier is radically different from the model (2.34), where the strong nonlinearity generates
power. harmonics of the carrier and beat notes.
According to this model, the I/ f noise arises from the amplifier's nonlinearity, and
it should thus be possible to reduce it by an improvement in linearity, Yet, most of Flicker noise in real amplifiers
the linearization methods known in the literature fail in this case because they act on It turns out that both mechanisms are present in real electronic devices, i.e. both non-
the entire amplifier, while the l / f noise is generated at a microscopic level inside the linearity and modulation of the gain, and both originate 1/f noise. In a deeper analysis,
electronic device. An example is the case of the push-pull amplifier. Because the even the difference between these two phenomena is more subtle than that presented here
harmonics are canceled by symmetry, one may expect a significant reduction in the because the nonlinear approach hides a parametric modulation. As a first approximation,
I / f noise, but this does not happen because the l / f noise is generated locally by the we can assume that in both models the statistical properties of the near-dc processes n'(t)
two transistors of the push-pull, and it is added statistically. Nonetheless a reduction of and nff(t)are not affected by the carrier power. This is confirmed by the experimental
3 dB in the l / f noise is expected, for reasons detailed below in subsection 2.5.4. Some observation that the amplifier phase noise given in r a d 2 / ~ zis roughly independent of
methods useful in reducing the I / f phase noise are explained in Section 2.5. the power in a wide range [45, 105,461. Some dependence on Po may remain in practice.
46 Phase noise in semiconductors and amplifiers 2.3 Phase noise and amplitude noise 47

Table 2.1 Typical phase flickering of amplifiers in dB rad2/Hz

GaAs HBT SiGe HBT Si bipolar


Rating (microwave) (microwave) (HF-UHF)

fair -100 - 120


good -1 10 -120 -130
best - 120 -130 - 140

We ascribe this to terms of order higher than vz in (2.3 1) and to the fact of operation in
a large-signal regime, which affects the dc-bias and in turn n'(t) and nJJ(t). fc i f
: - , i range of f,
In conclusion, the 1/f phase noise is best described by

s,( f ) = b-1 f -' (b-I % constant vs. 6 ) , (2.44)


Figure 2.7 Typical phase noise of an amplifier. Reproduced from [83] with the permission of the
IEEE, 2007.
where the coefficient bWlis an experimental parameter of the specific amplifier. The
same holds for the 11f amplitude noise. Table 2.1 shows the typical flicker phase noise
of commercial amplifiers. Phase flickering is related to the physical size of the amplifier's -120 1 ' ' I * I I ' I ' '
- SiGe LPNT32 -
active region and to the amplifier gain. The latter is a side effect of the number of stages I I bias 2V, 20 mA
needed for a given gain. More details are given in subsection 2.5.3.

2.3.3 Power spectral density and corner frequency


Let us combine the white and flicker noise, under the assumption that they are inde-
pendent. This is in agreement with observations on actual amplifiers. Thus we have for
S,( f ) (and see (2.30), (2.44))

I I I I I ' I
L..
1 a ' 1 1
1 10 102 103 104 105
f, Hz
Figure 2.8 Example of amplifier phase noise measured at various input power levels below
saturation. Courtesy of Rodolphe Boudot [I I].
bLl = constant.
Figure 2.7 shows the phase-noise spectrum of an amplifier. In this figure, we observe
that the corner frequency, where the white noise equals the flicker noise, is measured at the FEMTO-ST Institute in well controlled and understood conditions,
fits (2.45) well. The second amplifier was measured at liquid-He temperature at the
University of Western Australia [54, Fig. 21. The original data, given as the equivalent
noise temperature as a function of the input power, have been turned into the more
Recalling that bo = FlcT/Po, the above becomes readable form of Fig. 2.9. For input powers 0.1-10 pW (-70 to -50 dBm), the l /f
phase noise is fairly constant. The broken-line plot (-80 dB m) is clearly polluted by
the background noise of the instrument. In the reported experiment the white noise, not
As a trivial yet relevant conclusion, f , is proportional to the carrier power Po. shown in Fig. 2.9, was visible only at f > 10 kHz.

Experimental evidence Mistakes commonly found in amplifier specifications


Figures 2.8 and 2.9 show the phase-noise spectra of two microwave amplifiers measured Numerous data sheets give a specification for white phase noise, in dBc at a suitably
at different power levels below compression. The first amplifier (Fig. 2.8), which was large offset frequency, but often without indicating the amplifier's input power. Such
48 Phase noise in semiconductors and amplifiers 2.4 Phase noise in cascaded amplifiers 49

is still somewhat unclear, the l / f noise seems to be due to the propagation of heat in
= -50 ~ B I Amplifier
-60 dB m electronic circuits; this filters the temperature fluctuations of the environment. The 1/f
-70 dB m
slope has also been observed in precision dc amplifiers [95,86]. Better thermal shielding
-80 dB m
and higher thermal inertia shift the 1/f region towards lower frequencies. In reality, the
- -
= = a +P = -80dBm
l / f region is at most a large bump because the low-frequency asymptotic slope cannot
be steeper than - 1, otherwise the amplifier group delay would diverge rapidly. Another
-70 dB m type of bump has been reported [70], associated with positive feedback or "ringing" of
-60 dB m the heat propagation inside the amplifier.
-50 dB m

-130 4
Phase noise in cascaded amplifiers
I I

1 101 lo2 103 104 1o5


Fourier frequency, Hz
The phase noise of a cascade of amplifiers is governed by the rules indicated in Fig. 2.1 1
Figure 2.9 Phase noise of the amplifier X-9.0-20H measured at 4.2 I<, at different input-power
and detailed below. Although our discussion is in terms of the phase noise, the amplitude
levels. The experimental data are talten from [54, Fig. 21.
noise is governed by the same laws. It turns out that the rules for parametric noise are
at first sight counter-intuitive and radically different from any experience one may have
a specification is misleading because it hides the nature of the white noise, which is
had with white noise.
identified better from the noise figure or the noise temperature.
Another common mistake is to specify the flicker noise in terms of the corner fre-
quency f,.This has been seen in a number of technical articles and data sheets. Once White phase noise
again, the use of f, is misleading because fc depends on the power. Instead, phase In a chain of amplifiers, the white noise adds up at the input of each device, regardless
flicltering should be specified in terms of its b-1 coefficient. of the carrier power. Thus, we can add the noise power of each device as referred to
the input of the chain and calculate the noise figure F , as in subsection 2.2.2. Then we
2.3.4 Environment-originated noise evaluate the phase noise using ba = F1zTo/Po,(2.30), which holds for additive noise in
general regardless of the amplifier's internal structure and number of stages. Therefore,
The phase lag and gain of an amplifier are sensitive to the environment that surrounds the amplifier is governed by an adaptation of the Friis formula (2.22) to phase noise:
it, chiefly the supply voltage and the temperature. Of course, a fluctuating environment
generates amplitude noise and phase noise; this mechanism is a parametric modulation.
For small fluctuations, the amplifier can be assimilated to a linear time-invariant system
(/,"chain = +7
( ~ 1 (white noise) . (2.50)

(Section 1.5), which is best described by its response h(t) to the Dirac S(t) function As before, the first stage has the highest weight in the chain.
or equivalently by its frequency response H ( j w ) . Letting T(t) be the environment
temperature, or any other parameter, the output amplitude and phase are
Flicker phase noise
It was shown in subsection 2.3.2 that, owing to the parametric nature of fliclter, the
phase-noise spectrum Sq(f ) = b-l /f is independent of the carrier power. Thus when
where the symbol * is the convolution operator (see subsection 1.5.1). The impulse rn amplifiers are cascaded, each contributes its own phase noise b-l/ f . Assuming that
responses h,(t) and h,(t) contain the time lag between environment and effect. Unfor- the amplifiers are independent, the noise adds up statistically. Hence the noise power is
tunately, in most practical cases we do not have sufficient information to use (2.48) and the sum of the individual powers, so that we have
(2.49). We must rely on the measured amplitude or phase-noise spectra alone.
(b-l)chain= (bml)i (fliclter noise) .
Thermal effects
It is observed experimentally that temperature fluctuations give rise also to a spectrum It should be noted that the Friis formula (2.22) is incorrect in this case. Instead, the
',
proportional to l / f as shown in Fig. 2.10. The frequency scale of this phenomenon phase-noise contribution of each amplifier is determined by the noise parameter b&l,
depends on the electronics and on the mechanical assembly as well. Though the question independently of the position of the amplifier in the chain.
50 Phase noise in semiconductors and amplifiers 2.4 Phase noise in cascaded amplifiers 51

Additive noise
F=Fl+-
F2- 1
IA,I~
+ ...
p (Fl - l ) k G (F2- l)kG
e

Fliclter noise
S,=S (01+Sp2

' = a' 1+ a' 2


a (random amplitude) (random amplitude)

Environmentally originated noise


phase phase

transfer functions
environmental
fluctution (T)

Figure 2.11 Propagation of phase noise in a cascade of amplifiers.

Environment-originated phase noise


According to subsection 2.3.4, this type of noise results from an environmental quantity
T(t) that modulates the carrier through an appropriate transfer function h(t). The quantity
T(t) can be the temperature, the humidity, the supply voltage, etc. Let us consider a chain
of amplifiers located at the same place or in the same box, powered by the same source,
etc. All these amplifiers experience the same fluctuating quantity T(t), which gives rise
to a phase noise q(t) = T(t) * h(t). Hence, the total phase fluctuation at the output of a
chain of rn amplifiers is

q(t) = T(t) * h 1(t) + T(t) * h2(t) + a . + T(t) * hm(t) (2.52)

or, equivalently,
52 Phase noise in semiconductors and amplifiers 2.5 * bow-flicker amplifiers

microwave microwave power main


input RF output output
mixer amplifier mixer vl(t) v,(f)
.
delay CP2
vi (t> - - - - - - - - - - s>
e
vif(t> ~$1 vo(t>
e
Ze
RF IF Vb V IF RF
vo vo (-10dBor-70dB) (-lOdB)

local error
oscillator main
input

Zm

Figure 2.12 Transposed-gain amplifier.


CP3 delay Cp4 noise 1
7
~z,(t) error
amplifier

using Laplace transform^.^ The phase-noise spectrum is obtained as S,( f ) = 1 a(f)12, Figure 2.1 3 Feedfonvard amplifier, showing the main and error branches.
after replacing s by j 2 n f , taking the one-sided spectrum, and averaging. The problem
with this formal approach is that in virtually all practical cases we have insufficient Feedforward amplifiers
information. So we can only make calculations for the worst case, in which all phases
originating from a single fluctuating parameter add up, so that Feedfonvard and feedback amplifiers were studied around 1930 at the Bell Telephone
Laboratories [7, 81 as a means of correcting the distortion in vacuum-tube amplifiers.
Feedback has been used extensively since, while feedforward was nearly forgotten.
Recently, wireless engineers realized that feedforward can be used to correct the non-
linearity of power amplifiers of code-division multiple access (CDMA) networlts and
Otherwise, we have to rely on measurements. other telecommunications systems in which a high peak-to-average power ratio limits
the use of highly nonlinear high-efficiency amplifiers.
The scheme of a feedforward amplifier is shown in Fig. 2.13. A general treatment of
* Low-flicker amplifiers these amplifiers is available in [76]. Here we focus on the noise properties, which may
be derived from three asymptotic cases.
As a consequence of the Leeson effect, to be introduced in Chapter 3, the 1/f phase noise
of the sustaining amplifier is of paramount importance for oscillator stability. Therefore, 1. In the first case we assume that the whole system is free from noise. Under these
conditions, we see that the delay t, and the loss t, (where the subscript refers to the
a number of techniques have been developed to improve this parameter.
main branch) must be set for the error signal ve(t) to be equal to zero. For reasons not
analyzed here, matching the amplitude and phase of the two paths is not sufficient: it
Transposed-gain amplifiers is also necessary to match the group delay [92].
2. In the second case, we add the noise [n,(t)lpa originating at the input of the power
The transposed-gain amplifier (Fig. 2.12) originates from the fact that the l / f phase noise
amplifier lteeping the error amplifier noise-free. After combining the output of the
of microwave amplifiers is higher than that of microwave mixers and of HF-VHF bipolar
power amplifier with the main input, the result is an error signal [ne(t)lpaat the input
amplifiers. To overcome this, the input signal is down-converted from the microwave
of the error amplifier. This error signal is amplified and added with an appropriate
frequency vo to a suitable frequency vb = vo - vr, amplified, and up-converted back to
weight to the delayed output of the power amplifier to form the signal [v2(t)Ipa
vo. The delay line, which matches the group delay t, of the amplifier, is necessary in
that nulls the effect of [n,(t)lp, at the main output. This second case illustrates the
order to cancel the local oscillator's phase noise, which is common to and coherent at the
noise-cancellation mechanism of the feedforward amplifier.
down-converting and the up-converting mixer. This type of amplifier was proposed as
3. In the third case, we consider a noise-free power amplifier and introduce the noise
the sustaining amplifier of microwave oscillators independently by Driscoll and Weinert
[ne(t)],, originating at the input of the error amplifier. Here, we note that [n,(t)lea is
[27] and by Everard and Page-Jones [3 11.
amplified and sent to the main output with no compensation mechanism.

The Laplace transform is denoted by the upper case letter corresponding to the time-domain quantity, except The above reasoning should indicate that the noise [n,(t)lpa originating in the power
for the pair T ( s ) t, T(t). amplifier is canceled within the accuracy limit of the feedfonvard path, while the noise
54 Phase noise in sernicsndu~torsand amplifiers 2.5 * Low-flicker amplifiers 55

ne(t)leaoriginating in the error amplifier is not canceled. Interestingly, the nature of the main
output
amplifier
noise n,(t)lea is still unspecified. This fact has the following consequences.
phase
mod
White noise input
Letting [ne(t)lea= FeakG be the white noise of the error amplifier, we find that the vi(0
P' RF& phase
equivalent noise of the feedforward amplifier is the noise of the error amplifier referred *
CP 1 / detector
to the main input. Thus, the noise figure is auadrature IIF
1 ' adjust. 1
error signal vd = kd (p2- pl)
F = Fea li (noise figure) , (2.56)
i

where ni lj is the product of all the power losses, dissipative and intrinsic, in the path
Figure 2.14 Noise-corrected amplifier.

from the main input to the input of the error amplifier. For this reason, it is good practice
Subsection 2.3.2 shows that, in the case of flicltering,the amplitude of the noise sidebands
to choose a low coupling factor for CP3 and CP4. For example, a - 10 dB coupler has
is proportional to the amplitude of the carrier (see (2.34)). Consequently, v,(t) and v,,(t)
an intrinsic3 loss of 0.46 dB in the main path. The white phase noise is bo = FkTo/Po,
can be made arbitrarily small by adjusting l , and t, to minimize V2. Imperfect setting
as in any other amplifier.
of l eor te9 as well as some residual carrier at the input of the error amplifier, results in
residual fliclter.
Distortion
In a well-balanced feedforward amplifier, the distortion of the power amplifier is fully
Oscillator application
corrected at the coupler CP2. Additionally, there is no carrier power at the input of
If the feedforward amplifier is overdriven,the power amplifier cannot provide the required
the error amplifier. Thus, the error amplifier amplifies only the distortion of the power
power. This results in a strong residual carrier at the input of the error amplifier and
amplifier, for it operates in the low-power regime. Under these conditions the error
ultimately in phase and amplitude flickering at the main output.
amplifier is fully linear, so the distortion compensation is nearly perfect. The residual
In oscillator applications, usually the gain control needed for the oscillation amplitude
distortion is due to an incorrect feedforward signal v2(t), i.e. to an incorrect l eor re,
to be constant relies upon saturation in the sustaining amplifier. In this case, the low-
and to a small nonlinearity of the error amplifier due to some residual carrier at its
fliclter feature of the feedforward amplifier, in comparison with traditional amplifiers
input.
may be reduced or lost. The obvious solution is to introduce a separate amplitude limiter
in the loop, provided that the 1/f noise of this limiter is sufficiently small. At microwave
Fliclcer noise frequencies this would seem to be viable using a Schottky-diode limiter because the
We have seen that the noise of the power amplifier can be completely removed by the l / f noise of these diodes is lower than that of microwave amplifiers. In optoelectronic
feedforward mechanism. Hence, in this paragraph we analyze only the l / f noise of the oscillators, it is possible to control the gain by exploiting the nonlinear characteristics
error amplifier. Let the signals added in the coupler CP2 be of the Mach-Zehnder optical modulator, so that the feedforward amplifier operates in a
vl(t) = Vl cos 2nvot , (2.57) fully linear regime.
Interestingly, the design of a feedforward amplifier for oscillator applications is sim-
v2(t) = V2 cos 2 n vot + v, (t) cos 2n vot - v,(t) sin 2n vot , (2.58)
pler than for the general case. The reason is that the signal bandwidth at the amplifier
where v,(t) and v,(t) are the in-phase and quadrature components of the error-amplifier input is narrow, limited by the resonator. Hence, it is sufficient to null the error-amplifier
noise, as in (1.9). For V2 << Vl, the amplitude and phase fluctuation at the main output input at the carrierfieqzkency instead of in a large bandwidth. Provided that the group
are delay is reasonably small, it is therefore sufficient to null the error-amplifier input by
adjusting the phase instead of the group delay.
vx (t)
a(t) = -, (2.59)
Vl
2.5.3 Feedback noise-degeneration amplifiers
The scheme of Fig. 2.14 maltes use of a low-noise phase detector that measures the
instantaneous phase fluctuation pe(t) = q2(t) - cgl(t) of the main amplifier by com-
The intrinsic loss of a coupler is due to energy conservation. paring the amplifier output with the input. The feedback control nulls the error signal
2.5 * Low-flicker amplifiers 57

m
main main
amplifier output

main
input
v,(0 input output

I
e--

bridge phase detector Vi (t)


---- ------------------- C--
I
I
I a
I
I
CP3
I
I
I
; A 2 I
I (noise) (carrier) (pump)
I
I
I
I
4
control //phase I

Figure 2.1 6 Parallel amplifier.

The main virtue of the bridge phase detector is its low l / f noise, which is limited by
the variable attenuator and variable phase shifter that balance the bridge. Flicker in the
error amplifier is virtually absent because the carrier power at its input is close to zero,
and parametric up-conversion of the near-dc flicker (subsection 2.3.2) does not take
Figure 2.15 Noise-corrected amplifier with bridge phase detector. The phase and amplitude
controls needed to stabilize the null at the A port are not shown.
place. The detector's white noise is bo = FelzTo/Pe, i.e. the equivalent noise FekTo of
the error amplifier referred to the power Pe at the input of the error amplifier. The power
and thus compensates for the phase noise of the main amplifier and phase modulator. loss in the path from the main input to the input of the error amplifier is the product
The control is effective for a bandwidth that can be of order 10-200 l a z . In the control ni l?of all the power losses l ? , dissipative and intrinsic. Thus, using the power Po at
bandwidth, the residual phase noise is chiefly due to the phase detector, including the the main input, the white phase noise is
output preamplifier (not shown). Using a double-balanced mixer as the phase detector,
this scheme is effective at microwave frequencies, where the mixer's lif' noise is lower
than that of the amplifier. Conversely, improving a HF-VHF amplifier in this way can
be more difficult because the l / f noise of amplifiers is already low. The noise of the corrected amplifier is determined by the phase detector in the
Reference [27] describes a microwave oscillator that makes use of a sustaining ampli- bandwidth of the control, and by the noise of the phase modulator and the main amplifier
fier of the type shown in Fig. 2.15, with a double-balanced mixer as the phase detector. outside this bandwidth. To minimize the flicker, the control bandwidth must be wider
than the flicker comer of the main amplifier. The couplers CP1 and CP3 impact on the
white noise.
Bridge implementation
The amplifier shown in Fig. 2.15 is a variation on that of Fig. 2.14 in which the double-
balanced mixer is replaced by a sophisticated bridge detector. The bridge is balanced Example 2.3. Let us assume that the noise figure of both amplifiers is 1 dB, that the
by setting the phase and attenuation in the CP4-CP3 path for the carrier to null at the loss of CP1 and CP3 is 3.5 dB (intrinsic plus dissipative), and that the noise figure of
A port of CP3. Hence, all the carrier goes into C, while only the noise is present in A. the phase modulator is 1.5 dB, due to a 1.5 dB loss. Under these conditions, the noise
The noise is amplified and down-converted to dc by synchronous detection. The variable figure of the corrected amplifier is 8 dB (due to the error amplifier and two couplers)
phase at the mixer LO port is set for the detection of phase noise. All details of this in the control bandwidth, and 6 dB outside the control bandwidth (one coupler, phase
low-noise detection method are explained in [82]. modulator, and the main amplifier).
In practice, this bridge requires that the phase and attenuation are automatically
controlled for the carrier null to be stable in the long run. This control is not shown in
Fig. 2.15. When inserting the amplifier into an oscillator, only the amplitude is to be Parallel amplifiers
controlled; the phase is set to zero by the oscillation frequency, which follows the drift A parallel amplifier has nz equal amplifiers connected in parallel, as shown in
of the bridge. An oscillator application of this type of amplifier is found in [55]. Fig. 2.16. The power splitter and the power combiner are assumed ideal, that is,
Phase noise in semiconductors and amplifiers 2.5 * Low-flicker amplifiers 59

impedance-matched, symmetrical, and free from dissipative loss. The parallel configu- Nonlinear model
ration extends the power output of a given technology because each branch contributes Introducing the flicker noise, the signal at the output of branch k is
l/m of the output power. Additionally, there are popular configurations that improve the
impedance matching by exploiting the properties of the 90" power-divider and power
combiner. Interestingly, the push-pull amplifier worlts as a parallel amplifier in which
two branches are combined in a 180" phase relationship. The above equation is (2.34) adapted to the symbols of Fig. 2.16. The random phase
Let us first analyze the white phase noise S,( f ) = bo.Denoting the input power by Po,
the power at the input of each branch is Po/m. Consequently, the phase-noise spectrum
is (bO)branch = mFlzTo/ Po. Combining the output of m independent branches, the phase
noise is at the output of each branch has a spectrum S$( f ) = [b-l]branch/
f , with
Flc To
bo = ------
Po
(parallel amplifier) . (2.62) [b- I] bra,ch = 4 4
a2
a l
S1j1/( f ) -

In summary, the advantage of combining rn independent noises is lost because the input Each p!rl,(t) contributes to the output phase p(t) by an amount
power is divided by m and hence the parallel configuration does not reduce the white
phase noise. Moreover, actual power dividers show dissipative losses, and this maltes
the white noise of the parallel amplifier higher than the noise of a branch.
In the case of fliclter noise, the coefficient bel of each branch is not increased by
dividing the input power by m. Hence, we expect a noise reduction. The analytical proof
is detailed below,
By inspection of Fig. 2.16, recalling that the power splitter and combiner are Thus the contribution of ql,(t) to the output phase spectrum is
impedance-matched, symmetrical, and free from dissipative loss, we find that
1
ulc(t) = -vi(t) (branch-amplifier input) , (2.63)
fi Notice that the expression (2.70) for pk(t) has as its denominator the total output
amplitude; thus plc(t) is the contribution of the branch k to the total output signal. The
and
output phase noise is found by adding the contribution of the nz branches:
1
vo(t) = - vlc(t) (main output) .
f i ,=I
Letting
Therefore we have
vi(t) = V,ejwot (input carrier) (2.65)

be the input carrier, we find zi,,(t) = (l/fi)V,ejwot at the input of each branch, and
vlc(t) = (al / f i V;,ejwot
) at the output. The coefficient a1 is the gain, i.e. the coefficient
under the assumption that the noise of all branches is equal (Sf,;(f ) = ST,//(
f)). Adding
of the first term in the nonlinear input-output function (2.3 1). Combining the m carriers,
the phase-noise spectral densities is a valid mathematical operation because the branches
we find
are statistically independent and because the spectral density is an average square quan-
vo(t) = a1 ~ , e j @ ~ ' (output carrier) (2.66) tity. Comparing (2.69) and (2.74), the fliclter coefficient of the parallel amplifier must
be given by
at the main output.
The proof of the l / f noise reduction is detailed for the nonlinear model of sub-
section 2.3.2 and briefly summarized for the quasi-linear model. The mathematical
manipulations and the results are the same. This proves the fliclter reduction property of the parallel configuration.
60 Phase noise in semiconductors and amplifiers 2.5 * bow-flicker amplifiers 61

Table 2.2 Specification of low phase-noise amplifiers. The four columns on the right give the phase noise at lo2, 103,
104,and 1O5 Hz. From AML Inc. [I]

Parameters Phase noise vs. f , Hz

Amplifier gain F bias power 102 1o3 1o4 1o5

AML812PNA0901 10 6.0 100 9 -145.0 -150.0 -158.0 - 159.0


AML8 12PNB0801 9 6.5 200 11 -147.5 -152.5 -160.5 -161.5
AML8 12PNC0801 8 6.5 400 13 -150.0 -155.0 -163.0 -164.0 A M T 813PNn11811l 181111 r n A \ fl
AML8 12PND0801 8 6.5 800 15 -152.5 -157.5 -165.5 -166.5
unit dB dB rnA dBm dB rad2/Hz -170 / ,
lo1 lo2 103 104 105 106
Fourier frequency, Hz
Quasi-linear parametric model
Figure 2.17 Phase noise of a family of AML amplifiers. From AML Communications Inc. [I].
In the case of the quasi-linear parametric model, (2.67) becomes

2.5.5 The effect of semiconductor size


Phase flickering is proportional to the inverse physical size of the amplifier's active
which is (2.39) adapted to the symbols of Fig. 2.16. Hence it holds that $/c(t) = nz(t),
region. This can be proved through a gedankenexperiment, in which we join the nz
thus [b-l]branch= Snll(f). After repeating the same mathematical manipulations as for branches of a parallel amplifier to form a large compound device. An additional hypoth-
the nonlinear model, (2.74) becomes
esis is required, that the near-dc flicltering taltes place at the inicroscopic scale, for there
is no correlation4between the different regions of the compound device. This hypothesis
is consistent with the two most accredited models for fliclter noise [49, 691.

which yields (2.75). This confirms the fliclter reduction.


2.5.6 SiGe amplifiers
Commercial amplifiers As a matter of fact, well supported by experience, the close-in 1/f noise is significantly
At the time of writing, AML Communications Inc. [I] has been identified as the sole lower in bipolar transistors than in other types of transistor. This fact is related to their
manufacturer of parallel amplifiers intended for low-phase-noise applications. Actually, superior linearity and to the greater volume of the region in which amplification taltes
the fact that the low phase noise is achieved by paralleling a number of stages is not place. Unfortunately, the gain of Si bipolar transistors drops at microwave frequencies,
stated, but the specifications (Table 2.2) leave little doubt that the number of branches beyond some 2-3 GHz. This limit has been overcome by recent silicon-germanium
connected in parallel scales in powers of 2. Looking at Table 2.2 from top to bottom, technology. A SiGe heterojunction bipolar transistor (SiGe HBT) is similar to a conven-
at every row the bias current is doubled and the maximum output power increases by 2 tional Si bipolar transistor except that the base is made of polycrystalline SiGe, which
dB. That the power increase is of 2 dB instead of 3 dB is sound, if one accounts for the has a narrower bandgap than Si. For detailed information about SiGe transistors refer to
loss and for the asymmetry of the internal passive networks. The gain reduction of about the boolts [2,23,22].
1 dB at every step reinforces this conclusion. The low l / f noise of SiGe HBTs is of interest in connection with microwave os-
The phase-noise specifications of the AML amplifiers are somewhat unclear because cillators, where it has been used successfully to implement a variety of high-spectral-
the l / f and the white region (Fig. 2.17) are not clearly identified. This could be due purity units spanning from commercial VCOs to research-grade sapphire oscillators
to a number of reasons, at which we can only guess, such as the use of conservative [12,20,41, 651.
specifications and production averaging. Nevertheless, some mismatch between the
model and the specifications does not change the main fact, that the phase noise decreases
by 2.5 dB every time the number of branches is doubled. The missing 0.5 dB may be Of course, this cannot hold at any scale. We expect that there exists a correlation length characteristic of the
ascribed to asymmetry in the power splitters and combiners. lattice or of the defect type.
supply 2.6.1 Modulation index
1 "cc For a given maximum peak power of the laser in Fig. 2.18, the highest microwave
power at frequency oo is obtained using square-wave modulation. For our purposes, this
source
r - - - _ _ _ _ _ - - _ _ _

I laser I
photodiode condition is equivalent to nz = 4/n 2r. 1.273, which is the first tern1 of the Fourier series
-----------
of the square wave switching between f1.
A case of great practical interest is that of the electro-optic modulator (ECBM), the type
most used in photonic delay-line oscillators (Chapter 5). The EOM optical transmission,
as a function of the driving voltage v9 is
; bypass;
I
~~iicrowave amplifier
input
V, being the half-wave voltage of the modulator. Driving the modulator with a microwave
signal v(t) = Vocos mot, the instantaneous transmission is
Figure 2.18 Detection of a ~~~icrowave-modulated laser bean1 ill sinusoidal or distorted regime.
Other types of source produce siiililar intensity modulatiols,

s Det~ctionof mi~oewsve-moduIat~d
light where J1 is the first-order Bessel function of the first kind. Equation (2.82) derives from
the k = 0 tern of the series expansion
4n this section we st~ldythe noise in the detection of a microwave-modulated laser beam.
This topic is of paramount importance for optoelectronic oscillators. We will refer to
the sinusoidal or distorted modulation regime (Fig. 2.18). With this restriction, most
sin(z cos 0) = 2 x(-
00

k=O
I ) ~ J ~ cos
~ +(2k
I + I)$ . (2.83)

of the microwave power is in the output bandwidth of the photodiode, so that we can
talte the hndamental output frequency. It should be pointed out that in the sharp-pulse The discarded terms "-. . " in (2.82) are harmonics of frequency noo, with i~itegern 2 2.
regirvle the photodetector operation changes, being dominated by the transient inside Though neglected fur the output signal at oo,their presence is essential in that they ensure
the semiconductor-.Although it is important in modern optics, this topic is beyond our that0 5 75 1.
scope. Comparing (2.82) with (2.781, the modulation index is
The instantaneous power Pl(t) of the optical signal sinusoidaliy modulated with a
modulation index 112 is

The maximu~nof (2.84) is m E 1.164, which occurs at Vo = 0.586 V,.


in this section, we use the subscripts I for "light9' and ""09 for "microwave." Equa- Harmonic distortion could be avoided by lteeping m small. Yet, this is not beneficial
tion (2.78) is formally sirnilar to the traditional amplitude modulation of radio broad- because harmonic distortion has no first-order effect on. noise and because the microwave
casting, but here optical power is modulated rather than 'ip9: voltage. Thus, if the in- harmonics are easily filtered out. However, increasing In increases the microwave power
stantaneous power is described by (2.78) the light8 electric field is modulated in both and thus the signal-to-noise ratio. The optical power is limited by saturation in the pho-
amplitude and phase. todetector. In practice, the microwave power and the dG bias of the E 8 M are sometimes
It follov;ls fiom (2.23) that the detector photocurrent is dificult to set and maintain at the anaximum ansdulation index. This is due to t l ~ epossi-
bility of bias drift and to the thermal sensitivity of the lithium aliobate out of which the
EOM is made.
Only the ac term 112 cos mot of (2.79) contributes to-
the microwave signal. The microwave
power fed illto the load resistance Ro is Po= Roi&,hence 2.6.2 Phase noise
The discrete nature of photons leads to shot noise of power spectral density Si(J') =
2.q; (A~/HZ)at the detector output. It follows from (2.25) that the spectrum of the
Phase noise in semiconductors and amplifiers

2.7 A 1 nW carrier is amplified by a cascade of the four different amplifiers A-D. What
are the orders of the amplifiers that malce the noise figure a minimum or a maximum?
For the two cases, calculate the noise figure. Also calculate and slcetch the phase-noise
spectrum S,( f ).
2.8 Two amplifiers C are connected in parallel with ideal 90" loss-free power dividers
and combiners. Calculate and sketch the phase-noise spectrum S,( f ) for input powers
1 mW, 2 mW, and 5 mW. What happens if the loss-free power dividers and combiners
are replaced by real ones having dissipative loss 0.5 dB?
2.9 Four amplifiers C are connected in parallel with ideal two-way 90" power dividers
and combiners. This means that there are three power dividers at the input and three This chapter introduces the basic noise phenomena in feedback oscillators. The first
power combiners at the output. Sketch the scheme. Calculate and slcetch the phase-noise phenomenon is the Leeson effect, by which the phase noise of the sustaining amplifier -
spectrum S,( f ) for input powers 1 mW, 2 mW, and 5 mW. What happens if the loss-free the amplifier that compensates for the resonator insertion loss and ensures stationary
power dividers and combiners are replaced by real ones having dissipative loss 0.5 dB? oscillation - is turned into frequency noise, causing the phase noise to diverge in the
long run. Second, the oscillator is followed by an output buffer, which contributes its
2.1 0 A high-speed photodetector of quantum efficiency q = 0.8 receives a 5 mW light own phase noise. Finally, the oscillator traclcs the frequency of the internal resonator as
beam at 1.32 pm wavelength. The detector output is.connected to a 50 S2 resistor. well as its frequency fluctuations.
Calculate the current flowing in the resistor and the shot noise (in WIHz). Calculate an
equivalent noise temperature that accounts for both thermal noise and shot noise.

2.1 1 In the setup of the previous exercise, the light beam is now intensity modulated
3.1 Oscillator fundamentals
by a microwave signal, so that the power swings sinusoidally between 4 mW and 6 mW.
The basic feedback oscillator (Fig. 3.1) is a loop around which the gain A of the sustaining
Calculate the microwave power in the resistor and the white phase noise.
amplifier compensates for the resonator loss at a given frequency frequency wo. For the
2.12 As a consequence of temperature fluctuations, the phase-noise spectrum of an signal to be periodic, i.e. exactly replicated after one trip around the loop, it is necessary
amplifier has a term b-s/f5 with b-s = r a d 2 / ~ z What
. happens if two such that the loop phase arg Ab(jw) be zero or a multiple of 2n at w = wo. The amplitude
amplifiersare cascaded or connected in parallel? And with three such amplifiers cascaded and phase conditions for stationary oscillation of a complex signal Ap(jw) are lmown
or in parallel? collectively as the Barlchausen condition:

IAP(jw)l = 1 ,
arg Ap(jw) = 0 ,

or equivalently

Ap(j w) = 1 (Barkhausen condition) . (3.3)

The unused input (0 V) on the left-hand side of Fig. 3.1 serves the purpose of mathe-
matical analysis and physical insight. This input is used to set the initial conditions from
which oscillation starts and to introduce the equivalent noise of the loop components.
In most practical cases, the resonator frequency response lj3(jw)12 has a sharp peak
at o = wo. It is therefore convenient to assume that the amplifier gain A is independent
of frequency: the small dependence of A on frequency, if any, is moved to /3(jw). The
product of functions Ap(jw), which are still unspecified, is shown in Fig. 3.1. For the
salce of clarity, we will leave until later the more general case of oscillation at a frequency
wo not equal to the resonator's natural frequency w,.
Heuristic approach to the Leeson eflect 3.1 Oseillat~rfundamentals 69

= q, a period is reproduced
after a round trip
co # coo each round trip attenuates I
the signal coo = 0,

Figure 3.1 Basic feedback oscillator oscillating at the exact natural frequency of the resonator. Figure 3.3 Bipolar and field-effect transistors can be regarded as a current source controlled by
the input voltage.

amplifier resonator amplifier resonator small signal

FT>I * out ' I :


i< ;; , tuning range
Figure 3.2 Negative-resistance (quartz) oscillator. A ' for oo
1

arg A/3
m I

The model of Fig. 3.1 is general. It applies to a variety of cases and is not limited to
electrical and mechanical systems. A little effort may be necessary to identify A and
j3(jw), as follows.
In the case of a two-port microwave cavity connected to an amplifier in a closed loop,
matching the real oscillator to the model of Fig. 3.1 is trivial. A less trivial example is the Figure 3.4 Starting the oscillator.
negative-resistance oscillator shown in Fig. 3.2. In this case, the feedback function j3(jw)
is the resonator impedance Z(jw) = V(jw)/I(jw). The current I ( jw) is the input and load, we have to change the sign of the current and thus of the resistance. The condition
the voltage V(jw) is the output. The resonator impedance is a complex function of Aj3(jo) = 1 requires that G, > 0, which is equivalent to a negative resistance.
frequency that takes a real value (a resistance) at w = wo. A negative transconductance
G, (here the subscript stands for "mutual") is used to represent the amplifier. We can
relate the oscillator of Fig. 3.2 to the general scheme of Fig. 3.1 by observing that 3.1 .I Starting the oscillor
the controlled current generator is a transconductance amplifier that senses the voltage Oscillations start from noise or from a switch-on transient. In the spectrum of such
V across the resonator and delivers a current I = G, V. Common bipolar transistors random signals, only a small energy is initially present at wo. For the oscillation to grow
and field-effect transistors can be regarded as a current source controlled by the input to a desired amplitude, it is necessary that IAj3(jw)l > 1 at w = wo for small signals
voltage (Fig. 3.3), and they can be arranged in a scheme that implements a negative (Fig. 3.4). With I Aj3(jw)l > 1, the oscillation rises exponentially at a frequency wo
conductance. defined by arg A/3(jw) = 0. As the oscillation amplitude approaches the desired value
The generalization of the above concept is that the amplifier input and output signals an amplitude control mechanism (not shown in Fig. 3.4) reduces the loop gain, so that
can be either voltages or currents, which gives four basic feedback topologies, well the loop reaches the stationary condition Aj3(wo) = I. Two solutions are preferred in
known in analog electronics [9 1, Chapter 81. electronic oscillators.
The question of signs deserves some attention. In circuit theory it is generally agreed
that the current is positive when it enters a load or leaves a generator. This goes with AGC. The amplitude is stabilized by an automatic gain control (AGC) external to
the fact that a generator is intended to provide power, and a load to absorb power. the amplifier, which reduces the gain in proportion to the amplitude and with a
Accordingly, when we look at the controlled generator as if it were a resistor, which is a suitable time constant. This solution is preferred when low harmonic distortion
70 Heuristic approach to the Leeson ef.Fect 3.1 Oscillator fundamentals 71

v(t) Clipping S(W) The excess power goes small signal


C to higher frequencies \ large signal I AP I = 1

, 1 , W
.
,
.
. .
1 ,
j-+-+i tuning range
* I
I
Figure 3.5 When a signal is clipped, the excess power is pushed into harmonics of order higher , I
I
than 1. In this example the clipping is almost symmetric and thus the even harmonics are
strongly attenuated.

is an important feature, i.e. in variable-frequency oscillators where cleaning the --p--i-,


- W
output with a filter is impractical. Examples are the bridged-T and the Wien-bridge
oscillators used in audio-frequency distorsiometers.
natural
frequencyWn : I

Fast clipping.In precision RF and microwave oscillators, large-signal clipping of oscillation ob = w, + Aw


the sustaining amplifier is the preferred way to stabilize the amplitude. High-
frequency stability is achieved because clipping is a fast process and thus virtually Figure 3.6 Tuning the oscillation frequency by the insertion of a static phase.
free from phase fluctuations. Figure 3.5 shows the effect of saturation. When the
input amplitude exceeds the saturation level, the output signal is clipped. The After introducing $, the Barlchausen phase condition becomes
actual behavior can be a smooth compression rather than hard clipping. Further
increasing the input level, the gain at o o decreases and the excess power is pushed
into the harmonics, at frequencies that are multiples of wo. Quartz oscillators, in
which the amplitude is stabilized by amplifier saturation, can attain the remarlcable Hence, the loop oscillates at the frequency
stability of a few parts in 1o - ~ ,with a flicker spectrum in the region of 1 Hz [79].
In summary, it is important to understand that, for real-world oscillators:
for which
1. It is necessary that I A/?(jw)l z 1 for small signals, otherwise the oscillation will not
start.
arg AB(jw) = -$ . (3.6)
2. The condition lAB(jw)l = 1 results from large-signal gain saturation. This is nec-
essary for the oscillation amplitude to be stable and allows the amplitude of the For reference, I,? > 0 means that the loop leads; thus the oscillator is pulled to a frequency
oscillation to reach a steady state. higher than the exact resonance (Aw > 0).
3. The oscillation frequency is determined only by the phase condition arg AB(jw) = 0. Of course, it is necessary that the oscillator saturates. Accordingly, the maximum
This is necessary for the oscillation to be regenerated after a round trip. tuning range is the frequency range in which

I AP(jco)I >1 (small signal) (3.7)


3.1.2 Pulling the oscillation frequency
is guaranteed, so that the gain can be reduced by saturation. Out of this range, the energy
To explain how an oscillator can be tuned to the desired frequency, we need the conceptual stored in the loop decays exponentially and hence no oscillation is possible.
distinction between the oscillation frequency wo and the natural frequency onof the In the linear region (see the lower right-hand diagram in Fig. 3.6), it holds that
resonator. Of course, the former generally lies close to the latter. The mathematical
details are left until Chapter 4.
A simple way to adjust the oscillation frequency is to introduce into the loop a static
phase $, external to the resonator (Fig. 3.6). This method is often preferred at microwave If the resonator is a simple circuit governed by a second-order differential equation with
frequencies, where line stretchers are available. low damping factor (i.e. a large quality factor Q), in the vicinity of the natural frequency
72 Heuristic approach to the Leeson effect 3.2 The Leeson formula

quartz resonator
equivalent
___-__--____-__-___
I
electrode I I C0
I
I I

---PEP-; external Figure 3.8 The phase noise of the amplifier and of all other components of the loop is modeled as
(
I I I

I
motional parameters frequency adjustment a random phase $ at the input of the amplifier.
I - - - _ _ _ _ _ _ _ _ _ _ _ _ _ _ _

Figure 3.7 Typical tuning scheme for quartz oscillators.


that the oscillator responds to it with a frequency fluctuation

onit holds that

d
d o arg A/3(jw) = --
- 2Q (or$argA,f3(u)=-- with associated power spectral density
o n vn ;
2Q) (3.9)
thus the fractional frequency offset introduced by the static phase $ is
Ao Au $ Ao 1
- -- --
- - for - << - . (3.10)
oo vo 2Q mo 2Q The instantaneous output phase is
Alternative tuning method
A second tuning method consists of pulling the natural frequency of the resonator by
modifying the parameters of the resonator's differential equation. The adjustment circuit
is no longer distinct from the resonator. In this case, there is no need to introduce the The time-domain integration corresponds to a multiplication by l l j w = l l ( j 2 nf ) in
static phase $. This method is often used in quartz oscillators, where a variable capacitor the Fourier transform, thus into a multiplication by 1/(2nf)2 in the spectrum. The
is used to alter the resonator's natural frequency (Fig. 3.7). factor 2n in (3.15) cancels with the 2n in l l ( j 2 n f). Consequently, the oscillator's slow
phase-fluctuation spectrum is

3.2 The Leeson formula

Let us consider a loop oscillator in which the feedback circuit P(jw) is an ideal resonator For thefast components of $(t), i.e. those faster than the inverse of the relaxation time
with a large1 quality factor Q, free from frequency fluctuations (Fig. 3.8). The relaxation r , the resonator is a flywheel that steers the signal. Loosely speaking, the resonator does
time of the resonator is not respond to fast phase fluctuations; its output signal is a pure sinusoid. Accordingly,
the fluctuation +(t) goes through the amplifier and shows up at its output without being
fed back to the input. No noise regeneration takes place in this conditions, thus
Replacing the static phase I+$ of (3.8) by a time-varying phase +(t) results in an oscillator
output phase that is a function of time:
and

In the case of slowflt~ctziationsof $(t), slower than the inverse of the relaxation time
r , the phase $(t) can be treated as a quasi-static perturbation. Equation (3.10) tells us
By adding the effect of fast and slow fluctuations ((3.16) and (3.18)), we get the
' Strictly speaking, only Q 2 10 is necessary. Leeson formula relating the oscillator's output phase spectrum to the amplifier's phase
74 Heuristic approach to the Leeson efliect 3.3 The phase-noise spectrum of real oscillators 75

sustaining amplifier output buffer


-----------
-----------

I resonator -,
(2-4 stages)

I random
I fluct. 601, I
_ - _ _ - _ _ - _ _ _ _ _ _ _ I
Figure 3.9 The modulus squared of the Leeson-effect phase transfer function H(jf).
Figure 3.10 Block diagram of a real oscillator consisting of resonator, sustaining amplifier,
and buffer.
fluctuations:
The formula (3.19) was proposed by David B. Leeson 1631 as a model for short-
S+(f) (Leesonformula). (3.19) term frequency fluctuations and was initially intended to explain the phase noise of the
crystal-oscillator frequency-multiplier source for airborne Doppler radar applications.
The Leeson formula (3.19) can be rewritten as As will be shown in Chapter 6, for the range of the baseband frequency f used in radar
systems [64], it is appropriate to ascribe essentially all the noise to the amplifier.

3.3 The phase-noise spectrum of real oscillators


where

vo 1 The Leeson effect has been introduced in terms of a transfer function governing the
f ~ = % = % (Leeson frequency) propagation of the amplifier phase noise to the phase noise at the oscillator output. Now
we will analyze the phase noise of a real oscillator consisting of the three main bloclcs
is the Leeson frequency. By inspection of (3.19), it can be seen that the oscillator
shown in Fig. 3.10, the resonator, the sustaining anzpl$er, and the oztpztt bzIffer. The
behaves as afirst-orde~~filter
with a perfect integrator (a pole in the origin in the Laplace
last is necessary for output isolation and, ultimately, to prevent the external load affecting
transform representation) and a cutoff frequency fL (a zero on the real left-hand axis in
the oscillation frequency. Each block contributes its own noise to the oscillator output.
the complex plane). The theory of linear2 systems describes an oscillator in terms of its
Here, it is sufficient to account for the amplifier's white and flicker phase noise, because
phase-noise transfer function H ( jf); we have
the environmentally originated noise shows up only at very low Fourier frequencies

l ~ ( j f ) l=~
S d f1
=I + 1
(5) 2
(Fig. 3.9) .
(subsection 2.3.4 and Fig. 2. lo), where resonator instability turns out to be the dominant
phenomenon.

Some final remarks deserve attention. We have obtained the Leeson formula by fol-
lowing a heuristic approach based on physical insight. The rigorous proof, given in 3.3.1 Noisy amplifier and fluctuation-free resonator
Chapter 4, confirms this result. Equation (3.19), and equivalently (3.20), is a math- For a given noisy amplifier, the phase noise is white at higher frequencies and of the
ematical property of the oscillator loop and is independent of the physical origin of flicker type below the corner frequency f,.When such an amplifier is inserted into an
the phase fluctuations $(t), which are normally addressed as phase noise. The origin oscillator, the result for the oscillator signal, before buffering, is the two basic spectra
of phase noise is still unspecified, and the noise of the resonator has still not been seen inFig. 3.11.
accounted for.
Type-1 spectrum (fL > fc)
Real oscillators are inherently nonlinear. Nonetheless in most practical cases phase noise is a small pertur- This is the most frequently encountered spectrum, often found in microwave oscillators
bation, and a linear analysis of the amplitude and phase is correct. and in high-frequency piezoelectric oscillators (2 100 MHz), in which fL is made
76 Heuristic approach to the Leeson effect 3.3 The phase-noise spectrum of real oscillators 77

Table 3.1 The effect of the buffer on the output phase-noise


spectrum if all amplifiers employ the same technology

No. of Oscillator
buffer stages flicker

: white phase

Figure 3.11 With a flickering amplifier, the Leeson effect yields two types of spectrum. A
noise-free resonator and buffer are provisionally assumed. Adapted from [83] and used with the
permission of the IEEE, 2007.
loop, and the phase noise at the output of the loop is the phase noise of the sustaining
amplifier (see (3.17)). Consequently the sustaining amplifier and the buffer are modeled
as cascaded amplifiers, and the associated noise spectra are added according to the rules
high by the high oscillation frequency and the low quality factor Q. By inspecting
given in Section 2.4. To account for buffer noise, the noise spectra of Fig. 3.1 1 are
Fig. 3.1 1 (left) from the right-hand side to the left-hand side, we first encounter the white
replaced by those of Figure 3.12, as explained below.
phase noise bof O of the amplifier transferred to the oscillator output and then the white
frequency noise b-2 f -2 due to the Leeson effect on the amplifier white noise. At lower
frequencies, where the amplifier flickering shows up, the oscillator noise turns into the White phase noise
frequency-flicker type bW3f -3. The phase flickering (b-1 f -') is not observed at the In the white-noise region, the phase-noise spectrum at the loop output is bo = FkTo/Po,
output. (2.30). The buffer phase noise is governed by the same law but with a higher input
power in the denominator because of the sustaining amplifier. Therefore, the sustaining
Type-2 spectrum (fL < f,) amplifier has the highest weight in summing the phase noise spectra, as deduced from
This spectrum is found in RF (2.5-10 MHz) high-stability quartz oscillators and in (2.50). Under these conditions, with careful design the phase-noise contribution of the
cryogenic microwave oscillators. In both cases, the resonator exhibits a high quality buffer can be made negligible.
factor. Looking at Fig. 3.11 (right) again from the right-hand side to the left-hand side,
the amplifier's phase noise changes from white (bof O) to flicker (b-1 f -') at f = f,.
Phase flickering (b-1 f ;I) shows up in the frequency region from fr to f,.The Leeson Flicker phase noise
effect occurs only below fL, where the oscillator noise turns into frequency flickering In the case of phase flickering, the amplifier noise is roughly independent of the carrier
(b-3 f -3). The white frequency noise (b-2 f -2) is not observed at the output. power, hence (2.50) does not apply. Instead, the noise spectrum is the sum of the single
-' -'
Comparing the two spectra, there can be either the f or the f noise type. The spectra, as given by (2.51). While the sustaining amplifier often consists of a single
presence of both, sometimes observed, can be explained using a more sophisticated stage, two or more buffer stages may be needed for isolation. Because of the Leeson
model which accounts for the output buffer's noise contribution. effect, if a sophisticated low-flicker amplifier (Section 2.5) is affordable then it should
be used as the sustaining amplifier, not as the buffer. These design considerations yield
3.3.2 The effect of the output buffer the conclusion that the l / f noise of the output buffer is expected to be higher than that
of the sustaining amplifier.
Adding a buffer to the oscillator loop, we notice that the buffer and oscillator are If similar technology is employed, the fliclter coefficient b-l is about the same for
independent. Still neglecting the effect of the environment, the buffer's phase-noise each stage, whether it occurs in the sustaining amplifier or in the buffer. Hence, if there
spectrum contains only white and fliclter noise while the oscillator's spectrum becomes are n buffer stages plus one sustaining amplifier, the total 1/f noise spectrum observed
significantly higher at f < f L because of the Leeson effect. Thus, assuming that the +
at the output is n 1 times the 1/f noise spectruni of one amplifier (Table 3.1). In the
buffer and the sustaining amplifier make use of similar technology, the buffer noise absence of any other information, we can guess that an oscillator will have one sustaining
will show up only at f 2 fL. In the f > fL region, there is no phase feedback in the amplifier and three buffer stages, for the estimated flicker of the sustaining amplifier is
78 Heuristic approach to the Leeson effect 3.3 The phase-noise spectrum of real oscillators 79

justified only inside the loop, where the Leeson effect takes place. As a consequence,
r- -2 total noise r- - 7 total noise the larger l / f noise of the buffer is expected to show up. Interestingly, this is the only
The output buffer lf3 Low-flicker sustaining type of spectrum that contains both 1/f and l /f noise types.
am lifier (noise-corrected)
noise is not visible and'normal output buffer The larger buffer 1/f noise hides the fO to 1/f transition, which is the signature of the
Leeson effect. Of course, the Leeson frequency can still be estimated by extrapolating
the f-2 segment. The continuation of the f-2 line crosses the white phase noise at
f = fL.
The type-1B spectrum is observed in some exotic low-noise microwave oscillators
that make use of a cryogenic resonator, for maximizing Q, and a noise-degeneration
amplifier.

Type-2A spectrum (fL < f,)


The region between fL and f, is affected by buffer flickering, which is higher than
the l/f noise of the sustaining amplifier because of the higher number of stages.
-'
As a consequence, the corner where the f noise turns into f-3 noise is pushed to a
frequency lower than the true fL.This corner frequency has the same graphical signature
as the Leeson effect and is easily misinterpreted.
The type-2A spectrum can be found in high-stability 5-10 MHz quartz oscillators,
and in microwave cryogenic oscillators. It is the only spectrum in which there is a
clearly visible 1/f region of the spectrum and in which similar technology is employed
for sustaining the amplifier and buffer. In this case, we can infer the l /f noise of the
sustaining amplifier by subtracting 6 dB from the l / f output phase noise, under the
assumption that there are three buffer stages (Table 3.1).

Figure 3.12 Effect of the output buffer on some oscillator noise spectra.
Type-2B spectrum (fL < fc)
The type-2B spectrum differs from type 2A in that the noise-degeneration amplifier
one-quarter (-6 dB) of the total noise at the oscillator output, with an error of 1 dB if makes the l / f noise lower. Yet the graphical pattern is the same. The spectrum does
there are actually two or four buffers. not contain 1/f noise. The 1/f noise is the phase noise of the output buffer, which
hides the Leeson effect.
Type-IA spectrum (fL > fc) One may expect to find the type-2B spectrum in some sophisticated 5-10 MHz quartz
The sustaining amplifier and the buffer make use of the same technology, for they have oscillators, where fLis of order 1-10 Hz. However, noise-degeneration amplifiers are
similar flicker characteristics. Yet the Leeson effect, occurring at fL > f,,turns the not employed in this type of oscillator because the resonator noise is higher than the
white noise of the sustaining amplifier into 1/f noise for f c fL.It is seen in Fig. 3.12 1/f noise originating from the Leeson effect. This will be analyzed thoroughly in
that the Leeson effect makes the buffer flickering negligible. Hence, the insertion of a Chapter 6.
buffer at the output of the loop leaves the spectrum unaffected.
The type-lA spectrum is typical of everyday microwave oscillators such as dielectric
resonance oscillators (DROs) and the yttrium iron garnet (YIG) based oscillators. The effect of resonator noise
Thermal noise is inherent in the dissipative loss of the resonator. This effect is included
Type-1B spectrum (fL > fc) in the equivalent noise at the input of the amplifier, as explained in Section 2.2. Other
For maximum stability, the oscillator loop employs a noise-degeneration amplifier (sub- noise phenomena, always present and far more relevant to the oscillator's stability, are
section 2.5.3), which exhibits reduced flicker. Conversely, the buffer is a traditional the flickering and the random walk of the resonant frequency. The fractional-frequency
amplifier because the cost and complexity of the noise-degeneration amplifier are -'
spectral density S,(f)shows a term h - I f for the frequency flicker and a term h -2 f-2
Heuristic approach to the Leeson eMect 3.3 The phase-noise spectrum of real oscillators 81

fluctuating Zi, fluctuating ZOut


electronics electronics
I I
sustaining amplifier
out
*

/ - - - ,- -phase free
-------- - J
I -- -.
/ \
I 1
I I
I I
I I

1 %{dZin}4 dQ %{dZOut}+ d Q
:,O{dZin}-do,,
-----------
O{d~~~,}-+dw,~
- resonator - /'
+~Po'~)J I
I random fluct. don

Figure 3.14 Effect of the impedance fluctuation of the sustaining amplifier.

shows a corner point (f -3 to f - 2 ) that is easily mistaken for f, because it shows the
same graphical signature. Conversely, in the case of type-2 spectra the resonator l / f
noise may hide fL and cross the I/ f noise of the buffer and of the sustaining amplifier
(thin grey brolsen line). The spectrum has the same graphical signature as the Leeson
effect, i.e. the slope changes from f -3 to f at a corner point. Of course, this corner
frequency is not the Leeson frequency. This behavior is typical of high-Q HF quartz
oscillators (5-1 0 MHz). If the resonator 1/f noise is higher than shown, it may hide
both fe and fL. In this case, only one corner point is visible on the plot, where the
Figure 3.13 Effect of resonator frequency fluctuations on the oscillator noise.
resonator noise (f -3) crosses the white noise of the amplifier. This behavior is typical
for the frequency random walls; S,( f ) and S,,( f ) are related by (1.72): of VHF quartz oscillators.
In all cases, the l/f4 noise (due to frequency random walls) of the resonator is
v,'
Sp(f> = - SJdf) . (3.23) the dominant process at sufficiently low frequencies. At even lower frequencies other
f2 processes show up, such as frequency drift and aging.
-'
Thus the term h-1 f of the resonator frequency fluctuation yields a term b-3 f -3 in
the phase noise, while the term h-2 f -2 of the former yields a term b-4 f -4 in the latter.
Of course, the resonator noise is independent of anything else in the oscillator and thus
adds to the noise of the electronics, which includes the buffer and sustaining amplifier. 3.3.4 * Resonator-amplifier interaction
After introducing the resonator noise, the basic spectra of Figure 3.12 turn into those In the previous sections, the sustaining amplifier has been implicitly considered as a
of Fig. 3.13. The purpose of this figure is to show the graphical signature of all possible perfectly directional device that transfers the input signal to the output, at most intro-
cases. For this reason the resonator noise, shown as a bold solid line of slope l / f ducing a randomly fluctuating phase in the transfer but having no effect on the resonator
and l / f 4, is somewhat arbitrary. As indicated, it can lie anywhere within the light grey parameters. However, that things may not be this simple is suggested by the observation
regions. The main feature of Fig. 3.13 is the presence of the resonator 1/f noise, which that coupling the resonator to the oscillator inherently introduces dissipation. As a con-
can be higher or lower than the 11f noise due to the Leeson effect. This identifies the sequence, the resonator's quality factor is lowered by a factor of approximately 0.8-0.3.
resonator, or the sustaining amplifier noise througli the Leeson effect, as the main cause This is observed in a variety of resonators that include the quartz resonator, the mi-
of frequency flickering. crowave cavity, either dielectric-loaded or not, the whispering-gallery optical resonator,
In type- 1 spectra, the resonator 1/f 3 noise may hide f,,but not fL, and may cross the etc. Of course, if the real part of the amplifier's input or output impedance fluctuates then
l / f noise due to the Leeson effect (thick grey broken line). In this case, the spectrum the quality factor also fluctuates. This is shown in Fig. 3.14. Similarly, if the imaginary
Heuristic approach to the Leeson effect 3.4 Other lrypes of oscillator 83

part of the amplifier's input or output impedance fluctuates then the resonator's natural Feedback sustains oscillation
at any w at which argAP = 0
frequency also fluctuates, and so does the oscillation frequency.
ator
For example, in a 5 MHz quartz oscillator, the pulling capacitance can affect the Output A selector circuit (not shown)
fractional frequency by ~ O - ~ /This
~ F means
. that a fluctuation of 5 x 10-l9 F would ac- is needed to select the
oscillation frequency
count for a fractional-frequency fluctuation of 5 x 10-14. This value constitutes a record
stability for a quartz oscillator. For reference, in fundamental capacitance-metrology,
a resolution of 10-lo in the measurement of a 1 pF Thompson-Lampard standard is
definitely not out of reach, which means a resolution of F, The high resolution
of the metrology methods probably cannot be transposed to the measurement of the
sustaining amplifier, for two reasons. The first reason is that the Thompson-Lampard
standard is a four-terminal capacitor, while the sustaining-amplifier input or output is
a two-terminal network. The second reason is that the sustaining amplifier should be
measured in actual conditions, oscillating in a closed loop, which is clearly a problematic
requirement.
The noise mechanism due to the amplifier-resonator interaction differs from the
Figure 3.15 Basic delay-line oscillator.
amplifier's input-output phase noise in that the amplifier's fluctuating impedance enters
in the resonator dynamic parameters instead of in the feedback, thus it has no cutoff at
the offset frequency w ~However,
. this noise mechanism is not reported in the literature. thus
Therefore proving or disproving its relevance in precision oscillators is an open research
problem.

For slow phase fluctuations it holds that


Other types of oscillator

Delay-line oscillator
and therefore
A delay line (Fig. 3.15) in the feedback path can be the element that determines the
oscillation frequency, rather than a resonator. In the frequency domain, a delay t is
described by ,8(jw) = e-JUT.Thus, the loop can sustain the oscillation at any frequency
wl for which arg A,8(jw) = 0, that is, Conversely, this oscillator's response to fast frequency fluctuations is far more complex
than in the case of a resonator. In fact, the delay line is a wide-band device and thus it
2nl 1 does not stop the fast phase fluctuations. Chapter 5 is devoted to this topic.
wl = - or vl = - for integer 1 . (3.24)
t t

A selector circuit, not shown in Fig. 3.15, is therefore necessary for the selection of a
3.4.2 Frequency-locked oscillators
specific oscillation frequency wo .
In quasi-static conditions, the Leeson effect can still be derived from (3.8), repeated This type of oscillator (Fig. 3.16) consists of a voltage-controlled oscillator (VCO)
here: frequency-locked to a passive frequency reference, usually a resonator. In this subsection,
we denote by wo the oscillator frequency and by w, the resonator's natural frequency.
The error signal v, is proportional to the frequency error Aw = wo - Wn of the VCO,
not to the phase error. The reason is that the resonator's transfer function ,8(jo) turns
For a delay t,it holds that d[arg Ap(jw)] = -7. To this extent, the delay line is equivalent the frequency fluctuations into phase fluctuations. Therefore, the control is afiequerzcy-
to a resonator of resonant frequency vl and quality factor loclced loop (FLL), not to be mistaken for a phase-loclted loop (PLL). The FLL, less
well lmown than the PLL, is commonly used in atomic frequency standards and in
lasers.
84 Heuristic approach to the Leeson elFFect 3.4 Other types of oscillator 85

oscillator
output
passive
frequency
reference
random
phase
phase
detector
output
[TI;
L I

control

t
Figure 3.16 Discriminator-stabilized oscillator. Figure 3.17 Pound oscillator [77].Adapted from [85]and used with the permission of the IEEE,
2007.
The random phase +(t) of Fig. 3.16 represents the residual noise of the system,
originating in the lowest-power parts of the circuit. It can be the noise of the phase detector Once again we find the Leeson effect, i.e. the frequency-to-phase conversion of noise that
or of the preamplifier (not shown) that follows the detector. We analyze this scheme in shows up as a multiplication by 1/f in the left-hand region of the phase-noise spectrum.
quasi-static conditions, thus assuming that the random fluctuation +(t) is slower than the The complete Leeson formula (3.19), derived for the feedback oscillator, contains the
relaxation time of the frequency reference. The error signal is v, = k,[arg @(jw) +I, + white phase-noise term "1 ," dominant beyond fL, which is not present here. Something
where lz, is the gain of the phase detector. Using a resonator as the frequency reference, similar also happens in the case of the frequency-stabilized oscillator. Yet the present case
close to the resonance it holds that arg @ = -2Q(Aw/o,). Thus, the error signal is is more complex, because the gain of the phase detector can drop beyond fL and because
the VCO has its own white noise. Hence the structure of the frequency-stabilized oscilla-
tor must be detailed for a complete evaluation of its phase-noise spectrum to be possible.

The VCO is governed by the law o o = Iz,v, +


or,, where k, is the gain in rad/(V s), v, is
the control voltage, and ws is the free-running frequency at v, = 0. A dc voltage added 3.4.3 Pound stabilized oscillators
to v,, which may be needed to center the system variables in their dynamic range when
the error signal is zero, is ignored here because it has no impact on the noise. Denoting A popular implementation of the discriminator-stabilized oscillator is the Pound scheme
by k, the transfer function of the control, still unspecified, and introducing the VCO law [77], shown in Fig. 3.17. The modulation frequency f, is significantlyhigher than the res-
(see below (3.30)) and the loop gain kL = k,k,k, into (3.30), we find onator bandwidth v,/ Q, thus the modulation sidebands vo f f,, are completely reflected.
The carrier vo is partially reflected. The imaginary part of the reflected carrier is propor-
tional to the frequency error vo - v,, with a nearly linear law in a small interval. Com-
bining the sidebands and carrier in the power detector, which has a quadratic response,
thus an error signal of frequency f,, is present at the detector output. This signal, down-
converted to dc by synchronous detection, controls the VCO to null the error vo - v,
in the closed loop. The real part of the reflected carrier, governed by the impedance
mismatch at the resonator input, yields a dc signal, which is not detected.
Substituting A o = 2n Av and o, = 2n v, and assuming that k~ (2 Qlv,) >> 1 (large The main point of Pound stabilization is that the use of phase modulation moves
control gain), the VCO error (3.32) turns into the frequency-error information from near-dc to the modulation frequency f,, which
is farfiom thejliclcer region of tlze electronics. This feature reduces the Leeson effect
dramatically.
This result is the same as (3.13), derived for a feedback oscillator in quasi-static condi- Another advantage of the Pound scheme is that the lengths of the microwave cables
tions. Hence (from the VCO to the circulator, and from the circulator to the resonator) cancel in the
frequency-stabilization equations. As a relevant consequence, the lerzgtlz jlzictziations
impact only on the carrier phase, not on the frequency. In cryogenic oscillators, this fact
86 Heuristic approach to the Leeson efFect 4 Other Qpes of sseillator 87

laser
is0
phase optical makes it possible to use room-temperature electronics, far from the cooled resonator, at
modul. resonator a reasonably small loss of stability.
VCO
A version of the Pound scheme adapted to the stabilization of a laser frequency [26,6],
is shown in Fig. 3.18.
The frequency stabilization inherently requires a VCO locked to the passive reference.
It turns out that the stability of this VCO is a critical issue in high-demanding applications.
A smart solution is the Sulzer oscillator [98], shown in Fig. 3.19. In this scheme, the same
Figure 3.18 Pound-Drever-Hall laser frequency control [26].
resonator is used as the resonator of the VCO and as the passive frequency reference to
which the VCO is locked.
The Sulzer oscillator was invented to solve the problem of high fliclter in the early
transistors used in quartz oscillators. In fact, the frequency-stabilization feedback loop
f
compensates for the phase flickering of the sustaining amplifier and thus reduces the
I

I oscillator
I
I
Leeson effect. Of course, the frequency-error detection worlts at the audio-frequency
f,, far from the fliclter region of the electronics. However smart, Sulzer stabilization is
no longer used because the phase flickering of modern transistors is low enough to keep
the Leeson effect below the frequency fluctuations of the quartz resonator. This will be
analyzed thoroughly in Chapter 6.
The microwave version of the Sulzer oscillator, lcnown as the Pound-Galani oscillator
[34], is shown in Fig. 3.20. Of course, at microwave frequencies the Leeson effect is
still a major factor limiting oscillator stability. The Pound-Galani is one of the preferred
schemes for cryogenic oscillators because the VCO benefits from the high Q of the
cryogenic resonator.

,,
Figure 3.19 Sulzer oscillator [98].

--------.........................
oscillator + ;
I

I
I

'"0 ;
I
I
I output Q j
----
I /
I

--

Figure 3.20 Pound-Galani oscillator [34]. Adapted from [85] and used with the permission of the
IEEE, 2007.
4.1 Resonator d i ferential equation 89

input -
amplifier noise-free
I

oscillator
phase noise phase noise

'-LP resonator

Figure 4.1 Input-output phase-noise model of an oscillator.

The main purpose of this chapter is to prove and generalize the Leeson formula (3.19),
which was obtained with heuristic reasoning in Chapter 3. This extension in our laowl-
edge suggests new simulation and experimental techniques and enables the analysis of
other cases of interest not considered in the current literature, such as mode degeneracy
z~= sL 41 inductor
vL(t) = LL
dt i(t)

or quasi-degeneracy in resonators or in an oscillator pulled off the resonant frequency.


The analysis of delay-line oscillators and lasers in Chapter 5 is based on the ideas resistor
v(t) ZR = R
introduced here. vR(t)= R i(t)
Before tackling this proof, however, we must build up a set of tools to manipulate
the oscillator phase noise using Laplace transforms and the general formalism of linear
time-invariant systems. The underlying idea is to represent the oscillator as a noise-free
system that accepts a phase noise Q(s) at the input and delivers a phase noise @(s) at
the output, as shown in Fig. 4.1. In this way the oscillator may be described by its phase-
noise transfer function. The input noise, of course, is the noise of the oscillator's internal
parts. The use of a Laplace transform to analyze the phase fluctuation of a sinusoidal Figure 4.2 An R L C series resonator.
signal is inspired by the field of phase-locked loops (PLLs), where it is a common way
of calculating the transient response. However, this powerhl approach constitutes a new the fact that (4.1) originates from a positive-coefficient system, which in the case of the
departure in the noise analysis of oscillators. Consequently, a little patience is required electrical resonator means L > 0, C > 0, and R > 0, and to the physical need for the
as we go through a few sections of mathematics, which at the end will be gathered into resonator's internal energy to decay.
the Leeson formula. The symbols are inspired by the RL C series resonator, which is used as a canonical
example in this chapter. In this case, (4.1) follows immediately from Fig. 4.2 on removing
+ +
the generator, i.e. on setting vL(t) vR(t) vC(t) = 0, with the replacements
4.1 Resonator differential equation
w,2 = - (natural frequency w,)
LC
4.1 .I Homogeneous equation
and
A large variety of resonant systems are described by a second-order homogeneous
differential equation of the form Q = -w,L
1 (quality factor Q) .
--- -
R w,RC
d w, d
- i(t)
dt2
f - - i(t)
Q dt + wi i (t) = 0 (resonator) , Other more complex resonant systems can be approximated by (4.1) in the region around
w,. Distributed systems, such as the microwave cavity resonator and the Fabry-Pkrot
where i(t) is the current (or another variable that describes the specific systern), w, is etalon, are also locally approximated by (4.1).
the natural frequency, and Q is the quality factor. The conditions w, > 0 and Q > 0 We will search for the solutions of (4.1) in the time domain. The importance of this
are necessary for the resonator to be physically realizable. These conditions relate to approach is that no more than a minimum of knowledge about differential equations is
Phase noise and feedback theory 4.Hesonator differential equation 91

needed. We will guess a solution of the form1 theorem to s,:

i(t) = Ioest with s =o + jo , (4.4)

because the complex exponential eSt is an eigenfunction of the derivative operator, that
is, (dldt) e" = sseSt. After substituting i(t) = Ioestin (4.1),
= a: . (4.12)
The same holds for sp*because [91{s}12= [%{s*}12and [3{s}12= [3{s*)12.
we drop the time-dependent term Ioest. This is possible because (4.5) holds for any time Now we turn our attention back to the homogeneous equation (4.1). Since the co-
t . In this way, we obtain the associated algebraic equation efficients are real, the solutions must be real functions of time. Thus, using cosx =
+
f (ejx e-jx) and sins = -f j(e~*'- e-jx), the solutions of (4.1) take the form of a
damped oscillation:

whose solutions are


i(t) = d cos o p t e-'/' + 9sin o p t e-'li (solution of (4.1)) (4.13)
where d and 9are real constants determined by the initial conditions, and
2Q
t = --- (relaxation time) , (4.14)
a
11
The discriminant A, defined as
1

(free-decay pseudofrequency) . (4.15)

For high-quality-factor resonators, it holds that


nulls for Q = & f . For 1 Ql < f it holds that A > 0, hence the solutions of (4.6) are real.
The case Q < 0 is of no physical interest because it describes a resonator whose
internal energy builds rather than decaying.
up = on(1 - &) for Q >> 1 (4.16)

For 0 < Q < f , the solutions of (4.6) are real and negative. The solutions are real
because the discriminant is positive. They are negative by virtue of Descartes' rule of and so
signs,2 after observing that all the coefficients of (4.6) are positive.
For Q > f,which is always true for resonators of practical interest, it holds that A < 0.
By virtue of Descartes' rule of signs, the solutions of (4.6) are complex conjugate with
Equation (4.13) requires that t > 0 for the resonator energy to decay, which explains
a negative real part. After some rearrangement, the solutions of (4.6) are written as
why the condition Q > 0 is necessary. The relaxation time t is related to the other
resonator parameters by the following useful formulae:

t = - TQ n - - -Q= - =2Q
- --- 1 - 1 (relaxation time) , (4.18)
n nun on WL 27tf~
A relevant property of (4.6) is that the complex conjugate solutions are on a cir-
cle of radius on centered at the origin. This is easily seen by applying Pythagoras' where fLis the Leeson frequency.An example of damped resonator oscillations is shown
in Fig. 4.3.
In this section we denote the complex variable by s = 0 + j w because of the formal similarity to the results
obtained with the Laplace transform. In spite of this similarity, the variable s does not need to be identified 4.1.2 lnhomogeneous equation
with the Laplace complex frequency.
A real-coefficient polynomial of degree m has n z roots which are real or occur in complex conjugate pairs. Adding a forcing term v(t) to the series resonator of Fig. 4.2, the resonator is described
After arranging the polynomial in descending order of the variable and negating the coefficients of odd-
power terms, Descartes' rule of signs states that the number of negative roots is the same as the number of
by the inhomogeneous equation
sign changes in the coefficients, or a multiple of 2 less than this. In the case of complex conjugate roots, for
a pair of sign changes in the coefficients the polynon~ialhas a pair of complex conjugate roots whose real
part is negative.
92 Phase noise and feedback theory 4.2 Resonator Laplace transform

where the product in the numerator is taken over the N zeros and the product in the
denominator is over the M poles, and where C is a constant that is determined by
the residues. It may be remarked that the terms s - s, and s - sp are interpreted as the
distances of the points from the zeros s, and from the poles s,. This malies the evaluation
of I F(jo)12 a simple application of Pythagoras' theorem:

Using the R L C series resonator as an example, the admittance is

with

Figure 4.3 Resonator damped oscillation.

We restrict our attention to the case of sinusoidal forcing of frequency oo. Of course, the
+ +
This follows immediately from Z(s) = sL R l/sC, as seen in Fig. 4.2. In the
sinusoidal regime, for Q >> f the admittance has its maximum I Y(s)l = Q/wL = 1/R
asymptotic response is a sinusoid of the same frequency wo and of appropriate amplitude
at s = jw,.
and phase. This is necessarily so because for t + oo the energy contained in the initial
For the salse of generalization, we prefer a transfer function of the form
conditions is dissipated exponentially with a time constant equal to half the relaxation
time. Owing to linearity, the forcing response is added to the homogeneous solution. o n S

Hence the complete solution of (4.19) is of the form


P(s) = -
Q s 2 + o,s/Q + (general resonator transfer function) , (4.25)

which is obtained from (4.23) after multiplication by w,L/ Q. The function P(s) is
i(t) = d'cos wpt ,-'Ir + 28 sin o p t e-'IT + (e cos wot + sin mot , (4.20) interpreted as a dimensionless transfer. jiirzction normalized for lP(jw,) 1 = 1, as in
Chapter 3. With this choice, the Barlchausen condition for stationary oscillation is met
where d, 28, (e, and g are real constants and up is the free-decay angular pseudo-
with an amplifier of gain A = 1. Of course, an impedance or an admittance can still
frequency. For a given resonator, d' and 28' are the same coefficients as those of the
solution (4.13) of the homogeneous equation, determined by the initial conditions. be regarded as a transfer function, with the current as the input and the voltage as the
output or vice versa. A number of other resonant systems of interest are described, or
However, the coefficients (e and g are determined by the forcing term.
well approximated around the frequency con,by a function like (4.25). Figure 4.4 shows
the resonator transfer function in the frequency domain and the roots in the complex
plane.
Resonator Laplace transform Owing to the formal similarity of the denominator of P(s) to (4.6), the poles of P(s)
have the same properties as the solutions of (4.6), namely
In linear circuit theory, the use of the Laplace transform is common because it gives
access to a simple and powerful formalism for manipulating the network functions, i.e. 1. The poles are real and negative for 0 c Q 5 4.
the admittances and transfer functions. Using the Laplace transform, an R L C network is 2. The poles are complex conjugates for Q > f .
+
described by a rationalfunction of the complex variable s = o jw that is completely i,
3. For Q > the poles are on a circle of radius w, centered at the origin.
detemzinedby its roots (poles and zeros) on the complex plane. More precisely, a function
+ +
F(s) having N zeros s, = oz jw, and M poles s, = op jw, can be written as When the poles are complex conjugates, the system function P(s) is often rewritten
as
wn s
B(s) = Q (resonator poles s,, s; = 9A= jwp) (4.26)
(S - s~)(s- s;)
42 Resonator Laplace hansform 95

Co~npllewplane Wble 4.1 Relevant resonance parameters

frequency response

real part

% { p ( j w ) )= ---
ex ilnagiliiary part
1 + Q2x2
1
IP(j4l =
Jrn ~O~LIIUS

asg p ( j w ) = - arctan Qx argument

The ft~nctionP(s), as any network function, has the following properties:


Figuw 4.4 Resonator transfer fuilction P(s), (4.251,for Q >> 112. As usual in complex analysis, P(s) = P*(s") , (4.33)
the zeros are represented by circles and the poles by crosses on the coinplex plane.
/P(j@)l= lP(-ju)/ (the modulus is even),, (4.34)
with arg p(jw) = - arg p(- jw) (the argument or phase is odd), (4.35)

%{P(jw)) = W ( - j @ ) ) (the real part is even), (4.36)


WB(jw)) = -s{P(-jo)) (the imaginary part is odd) . (4.37)
These coi~ditionsare necessary for P(s) to be a real-coefficientanalytic function, or by ex-
tension a function ~vl~ose series expansion has real coefficients of s, and ultimately for the
inverse transforin (i.e. the time-donnain impulse response) to be a real function of time.
Figure 4.5 shows the symmetry in the case of a very high quality factor, Q >>> f .
The frequency response is found by substiming s = j w in P(s): The most re~narlcablefact is that the resonance shape is almost sytnmetrical, Thus the
symmetry properties (4.34)-(4.3'7) are also a local approxi~natioriaround the resonant
frequency w, and, of course, around -w,. Introducing the frequency offset 8 , it holds that

Defining the dissorznrzce x as lP(j(-wll - 6111 = l P ( j ( - ~ l l+ 6))l


(the modulus is even), (4.38)
= IP(j(an + 0
1

-
IP(j(wn - J))l
(definition of the dissonance x) ,
(3
K =- @11
-- (4.3 1)
1 w argp(j(-w, - 8)) 2 - argP(j(-o, + 6)) (the pl-nase is odd) ,
we find the elations ships shown in Table 4.1. In the vicinity of the natural frequency, the argP(j(w, - 6)) --argPCj(wn +8))
following approximation holds:
D?{P(j(-w, - 8))) = WP(j(-w, + 6))) (the real part is ever:), (4.40)
g{P(j(w,, - 8))) WBQj(wIl+ 8)))

Unfortunately, this approximation can be used only at positive frequencies. However, XP(j(-all - 81)) 2 -%P(j(-wIl i-
8)))
(the imaginary part is odd) . (4.41)
the negative-frequency properties can be obtairled by symmetry. XP(j(an - 8))) 2 -5(P(j(w, -t- 8)))
Phase noise and feedback theory 4.3 The oscillator 97

(a) Oscillator
-I small gain
output
Vo(s)

initial
conditions
Vi (s
I
-- ----- -.
(oscill. out)
v;<s>
or noise I 'I' I

(b) Classical control


large gain

Figure 4.6 Similarity and difference between an oscillator and a classical control.

The topology of the oscillator loop shown in Fig. 4.6(a) is similar to that of the
basic scheme used in classical control theory (Fig. 4.6(b)). The analogy enables a
straightforward derivation of the oscillator equation from control theory. From this
standpoint, an oscillator can be regarded as an ill-designed simple control, made to
oscillate by (intentionally) inappropriate feedbaclt. The main difference between the
Figure 4.5 Symmetry properties of the resonator transfer function /3(s).
oscillator and the control is that the oscillator has unity-gain positive feedbaclt instead
of large-gain negative feedbaclt; this is emphasized by the sign "+"
at the feedbaclt
These properties are a consequence of the proximity of the poles to the imaginary input of C . In this figure the oscillator output V, is talten at the amplifier input instead
axis. In this case, by virtue of (4.21), when the frequency axis is swept around the of at the amplifier output. The purpose of this seemingly weird choice is to simplify
resonant frequency o, only the distance from the pole close to jo, changes abruptly the equations. Of course, in actual implementations the oscillator output must be the
and symmetrically; the other roots of ,6(jo) are not much affected because they are far amplifier output V,/ in order to minimize the perturbation to the loop. Additionally, a real
away. The same thing happens when the frequency axis is swept around -on. By the resonator has an insertion loss, which is compensated by taking an amplifier gain larger
way, (4.2 1) also explains why arg ,6(jw) flips sign at o = 0, where P ( j o ) has a zero. than unity.
Referring to Fig. 4.6(a), elementary feedbaclt theory tells us that the oscillator transfer
function, defined as
The oscillator
VO(S)
H(s) = - (definition of H(s)) , (4.42)
Figure 4.6(a) shows an oscillator loop represented using a Laplace transform. The Us)
amplifier of gain A (constant) and the feedbaclt path P(s) are bloclts with which we are
familiar after Chapter 3. The signal F(s) at the input of the summing block C allows
initial conditions and noise to be introduced into the loop. Interestingly, it is also possible
to introduce a sinusoid of frequency close to the free-running frequency; this describes 1
H(s) = (see Fig. 4.6(a)) .
an injection-loclted oscillator. 1 - AB(s)
Phase noise and feedback theory 4.3 The oscillator 99

If the denominator 1 - Ap(s) nulls for s = fjwo, the system provides a finite response
(a) Oscillator transfer function N(s) (b) Detail of the denominator of N(s)
for a zero input, i.e. a stationary oscillation of frequency wo. From this standpoint, one
may regard the oscillator as a system with a pair of imaginary conjugate poles excited
by suitable initial conditions and interpret the frequency stability as the stability of the constant zeros
poles on the complex plane. Of course, the condition 1 - AP(s) = 0 for s = f j o o is
the Barlchausen condition introduced in Chapter 3.
i
, - g?$
,, .,

Mathematical properties
We first study the properties of H(s), (4.43), for the case when B(s) is the transfer
function of a simple resonator, (4.25), and the system oscillates exactly at the natural
frequency of the resonator:

Thus, substituting (4.25) into (4.43), we get


Figure 4.7 (a) Noise transfer function H(s), (4.45), and (b) root locus of the denominator of H ( s )
as a function of the gain A.

which can be rewritten as We first solve AD = 0, which yields

because a, = -w,/(~Q), (4.27). The above H(s) is a rational function with real coeffi- and hence
cients, for it can be written as N(s)/D(s), i.e. numerator/denominator where N(s) and
D(s) are second-degree polynomials. The function H(s) has two poles, either real or
complex conjugates depending on the gain A. The root locus is shown in Fig. 4.7. The
poles have the following properties. because op= -0,/(2Q), (4.27). As the coefficient of in AD is positive, we conclude
that
+
I. The poles are complex conjugates for 1 - 2 Q < A < 1 2 Q and are real elsewhere.
2. When the poles are complex conjugates, they lie on a circle of radius w, centered at
the origin.
3. The poles are imaginary conjugates for A = 1.

The proof is given below.


which constitute property 1 of the list preceding (4.47).
The poles of H(s) are the zeros of its denominator, i.e. the solutions of D(s) = 0:
When A D < 0, the solutions of D = 0 are complex conjugates:

This is an alternative form of (4.48), obtained after malting the square root real by
The poles are real or complex conjugates depending on the sign of the denominator changing the signs inside. The square distance R2 of the poles from the origin is
discriminant AD, defined as
100 Phase noise and feedback theory 4.4 Resonator in phase space 101

which simplifies to The gain control can be a separate feedback system that sets A for the oscillator output
voltage to be constant. This approach was implemented in early days of electronics by
the Wien bridge oscillator [48]. Amplifier saturation, however, proved to be an effective
This is property 2 of the list preceding (4.47). amplitude control even in ultra-stable oscillators. When the amplifier saturates, the
Finally, the poles are imaginary conjugates for A = 1. This is obtained by setting output signal is compressed more or less smoothly. The resultant power leakage from
A = 1 in (4.52). The discriminant the fundamental to the harmonics reduces the gain and in turn stabilizes the output
amplitude. The resonator prevents the harmonics from being fed back to the amplifier
input.
reduces to -w:, and the real part -an(A - 1) of the solutions sl, s 2 vanishes. Thus
sl, s:! = fjwn, which is property 3 of the above-mentioned list. Enhanced-quality-factorresonator
Positive feedback can be regarded as a trick to increase the quality factor Q of a resonator.
This can be seen by equating the denominator
4.3.2 Further remarks

About the amplifier gain


of the oscillator transfer function H(s), (4.43), to the denominator
Only the case A = 1 is relevant for oscillator design and operation. Nonetheless, the
analysis of other cases provides insight.

A < 0 corresponds to negative feedback. The resonator poles are pushed in the
of the transfer function Be&) of an equivalent resonator. The comparison gives
left-hand direction by the feedback, or made real for strong negative feedback
(A 5 1 - 2Q).
A = 0 corresponds to open-loop operation. The zeros and poles of H(s) cancel one
another, and the transfer function degenerates to H(s) = 1. This is consistent with thus, using ap= -wn/(2Q),
the choice of the output point (Fig. 4.6(a)).
0 < A < 1 corresponds to weak positive feedbaclc. The resonator poles are pulled
towards the imaginary axis without reaching it. The effect of the feedback is to
sharpen the resonator's frequency response, yet without stationary oscillation. The same conclusion can be drawn qualitatively by comparing Fig. 4.4 with Fig. 4.7,
A = 1 corresponds to part of the Barkhausen condition for stationary oscillation. The remembering that the complex conjugate poles are on a circle of radius w,.
poles are on the imaginary axis. Unfortunately, positive feedback cannot be taken as a way of escaping from the Leeson
A > 1 corresponds to strong positive feedback, which makes the oscillation ampli- effect by artificially increasing the quality factor. In fact the noise reduction achieved
tude diverge exponentially. The poles are on the right-hand half-plane, a > 0. For in this way vanishes because of the additional noise introduced by the amplifier used to
+
A > 1 2Q the positive feedback is so strong that the poles become real. In this increase the quality factor, unless a superior technology is available.
case, the output rises exponentially with no oscillation.

4.4 Resonator in phase space


Amplitude noise and frequency stability
That the poles of H(s) are on a circle centered at the origin has an important consequence
We will analyze the resonator phase response b(t) and its Laplace transform B(s) in
in metrology. In the real world the gainf-luctuates around the value A = 1. For small
quasi-stationary conditions. The resonator is driven by a sinusoidal signal at a frequency
fluctuations in A, the poles fluctuate perpendicularly to the imaginary axis. The effect
wo, which can be the resonator's natural frequency on or any other frequency in a
on the oscillation frequency is second order only.
reasonable interval around w,. In the time domain the phase transfer function b(t) is the
phase of the resonator's response to a Dirac 6(t) function in the phase of the input. More
Gain control precisely, b(t) is defined as follows (Fig. 4.8(a)). Let
The exact condition A = 1 cannot be ensured without a gain control mechanism. The
latter can be interpreted as a control that stabilizes the oscillator poles on the imaginary 1
vi(t) = 7cos(wot - 8 ) (stationary input) (4.57)
axis.
hase noise and feedback theor 4.4 Resonator in phase space 103

(a) Impulse response phase 0

phase impulse in cos[aot + d(t)] cos[u,ot + b (t)]


the voltage domain =-- resonator >

phase impulse in d(t) b(0


the phase domain equivalent >
) resonator

(b) Step response

phase step in cos[a0t + U(t)] cos[aot + bu(t)]


the voltage domain =- resonator .

phase step in u(t) I, b,(t) = j b(t) dt


the phase domain Figure 4.9 A sinusoid with a phase step can be decomposed into a sinusoid that is switched off at
resonator t = 0 plus a shifted sinusoid that is switched on at t = 0. The step response is the linear
superposition of the two responses.
Figure 4.8 Resonator phase response to a phase impulse 6(t).

Phase-step method
be a stationary dimensionless3 signal. The constants
The simplest way to find b(t) is to feed a small phase step K U ( ~into
) the argument of
the input signal instead of the impulse 6(t). This is shown in Fig. 4.8(b) for K = 1. The
function
00
0, t<O
are chosen so that the resonator's stationary output has unit amplitude and zero phase, U(t) = 6(t)dt = (Heaviside function) (4.63)
1, t>O
v,(t) = cos mot (stationary output) , (4.60) is the well-lmown Heaviside function, also called the unit-step function. The impulse
response b(t) is obtained from the step response bu(t) using the property of linear
when the stationary signal (4.57) is fed into the input. Then, introducing an impulse 6(t)
systerns that the impulse response is the derivative of the step response
in the argument of the input,
1
vi(t) = - cos[wot - 0 + 6(t)] (input impulse), (4.61)
Po That linearization is obtained for K --+ 0 is physically correct because in actual oscillators
we get the output transient the phase noise is a small signal.
The phase-step method, shown in Fig. 4.9, consists of decomposing the input sinusoid

which defines b(t). Of course, the impulse response b(t) is obtained after linearizing the
system. The resonator is described by a linear differential equation since the output is, into two truncated waveforms,
necessarily, a linear fbnction of the input. Yet understanding the process, and the meaning
of this linearization, requires some simple mathematics, which we now introduce.
+
vi(t) = ~ { ( t ) v;(t)

The system function does not depend on the physical dimension of the input. Thus, we use dimensionless
vj(t), switched off at t = 0 v?(t), switched on at t = 0
signals because this makes the formulae a little more concise.
Phase noise and feedback theory 4.4 Resonator in phase space 105

Y (t) = cos [mnt + K U(t)] We want to prove that the resonator impulse response in phase space is
generator 1
r----------------
I
~,(t)= cos [mot+ lcbu(t)] b(t) = -e-'I' (impulse response) (4.69)
; 1
I

I t=O t
cos ( o nt)
1
I I I
I
I 1 I
in resonator out $
=: WLe-W~t, (4.70)

That accomplished, the transfer function B(s) = C{b(t)) can be found in Laplace trans-
form tables to be
UL
B(s) = - (transfer function) (4.71)
s WL +

Figure 4.10 Electrical model of the phase-step method.


which is a low-pass function. Additionally, it holds that
so that the small phase step K U(t) can be introduced into the argument of v('(t) when
vr(t) starts. Here, the Heaviside function U(t) is used as a switch that turns from off to
on at t = 0, as shown in Fig. 4.10. Similarly, U(-t) switches from on to off at t = 0.
The output is In order to prove (4.69), we first observe that @(jwo)= 1; thus Do = 1 and 0 = 0.
The input signal is
v,(t) = v;(t) + vg(t) (starting at t = 0) , (4.67)
where the terms vA(t) and vt(t) are, respectively, the switch-on and switch-off transient
responses for t 2 0. The output signal for t c 0 has no meaning in impulse-response
analysis. As in this section wo and wn are talten to be equal, they may be used interchangeably.
Before going through the analysis of the resonator response, it should be pointed out The resonator response vA(t) to the sinusoid switched off at t = 0 is
that the phase-step method finds application in the following investigation approaches.
COS w,t, ti07
Analytic nzethod. This enables the calculation of the phase impulse response, as de- vo(t) =
coswpte-'IT, t>O,
tailed in the following subsections.
Sirnzllation tool. A driving sinusoid with a phase step can be regarded as the circuit where r = 2 Q/o, = 1/wL is the resonator's relaxation time and w, = on d m
shown in Fig. 4.10. This circuit is easy to simulate with any circuit-oriented is the free-decay pseudofrequency. Similarly, the response vt(t) to the switched-on
simulation program, the most popular of which is Spice. sinusoid is the exponentially growing sinusoid
Experimental technique. A phase step can be implemented with a phase modulator
driven by a low-frequency square wave. This can be a viable method for testing the
v:(t) = cos(w,t + ~ ) ( -1 e-'/') , t >O .
resonator inside an oscillator. For example, in a coupled optoelectronic oscillator For Q >> 1, we can malte the approximation wp 2 w, = wo. This is justified by the fact
[109], where the microwave resonance results from the interaction between a that the phase error ( accumulated during the relaxation time r is
microwave loop and an optical loop, the resonator cannot be separated from 1
( = (w, - wp)r = - .
the oscillator. Nonetheless, the phase-step method has proved to be useful for 4Q
measuring the closed-loop quality factor.
This is seen by substituting r = 2Q/w, and w, = w , d m into 5, and by
expanding in a series truncated at the first order for Q >> f .
Input signal tuned exactly to the resonator's natural frequency By virtue of linearity, the total output signal is
In this section we assume that the input frequency wo is tuned to the exact natural vo(t) = v;(t)+ vfi(t) , t > 0 ,
frequency w, of the resonator: = cos writ e-'I' + (cos w,t cos K - sin o,t sin K ) (1 - e-'Ir)
= cos writ (e-'I' + cos - cos ~ e - ~ / ' ) sin w,t sin K (1 - e-'IT) .
K -
Phase noise and feedback theory 4.4 Resonator in phase space

For K -+ 0 we use the approximations cos K 21 1 and sin^ 2: K .Thus Using the test signal

v,(t) = cos w,t - K sin w,t (1 - e-'IT) .


After factorizing out the time dependence w,t, the above can be seen as a slowly varying
phasor, switched off at time t = 0, the output is
/

cos wot , t ~ 0
vo(t) =
cosw,te-'1' t>O.
, normalized on K ,is the step response
The angle arctan (X{V,(~)]/~R{V,(~)}) For t 5 0 the output is determined by the choice of Po and 0, while for t > 0 the output
is free exponential decay that is independent of wo.
bu(t) = 1 - e-'IT (step response) .
Using the test signal
The derivative of bu(t) is the impulse response b(t) = (1lr)e-'l", which is (4.69). 1
vi(t) = - cos(wot - 0) U(t) ,
Po
* Detuned input signal switched on at time t = 0, the output is (see (4.20)),
Our aim is now to extend the results of the previous section to the general case where
the input frequency wo is not equal to the resonator's natural frequency w,. We want to
prove that the impulse response of the resonator is where d , %, V, and 9 are constants determined as follows. Given coo, the system has
four degrees of freedom: the amplitude and phase of vi(t), the resonant frequency w,,
b(t) = (Q sin Qt +5 eptl' (impulse response, Fig. 4.12) (4.74) and the quality factor Q. Thus the four unlcnowns d , $8, V, and g are completely
determined. After our choice of input amplitude and phase, the output for t + oo is
= (Q sin Qt + WL cos Qt) epwLt, (4.75) v,(t) = cos wot. This yields V = 1 and g = 0. Then d and 927 are found by using the
that the transfer function is continuity of the output signal at t = 0. This continuity condition gives d = -1 and
9 = 0. In summary,
1
B(s) = -
+
s l/z + a2t (transfer function, Fig. 4.13) ,
r (S + 1 / t -jQ)(s + l / t +jQ)
For Q >> 1, we make the approximation wp 2: w,. This is justified by the fact that
and that the phase error accumulated during the relaxation time, ( = (w, - wp)t = 1/(4Q), is
small for large Q. Consequently, from (4.8 1) and (4.83) the output transients are

v,(t) = cos w,t e-'li , t50 (switch-off) , (4.84)


v o ( t ) = - cos w,t e-" + cos mot t >0 (switch-on) . (4.85)

Focusing on the switch-on transient, we have the following input-output relationship:

where the frequency offset, or detuning, Q is defined as cos(wot - 0) U(t) ,


(4.86)
Q = wo - w, (definition of detuning, Q) . (4.79)
v,(t) = (- cos o,t e-'Ii + cos mot) U(t)
For reference, in the resonator bandwidth the detuning Q spans the interval - w ~to + w ~ . and, as an obvious extension,

Preliminaries sin(wot - 0) U(t) ,


Before introducing the phase step, we will study the transient of a resonator driven at
the frequency wo f: w,.
v,(t) = (- sin w,,t e-'1' + sin mot) U(t) .
Phase noise and feedback theory 4.4 Resonator in phase space

Introducing the phase step K at t = 0


We use the test signal (4.66), here repeated:

uf(t), switched off at t = 0 u!(t), switched on at t = 0

The input vf'(t) can be rewritten as


1
vy(t) = - [cos(wOt- (D) cos K - sin(w0t - (D) sin K] U(t)
Po
1
= - [cos(wot - (D) - K sin(u0t - (D)]U(t) for K << 1 .
Bo
Using (4.84) and the pairs (4.86) and (4.87) we get for the output signal

+
vo(t) = vL(t) v:(t) for t > 0 , ".V

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0


= cos ~ ,e-'/'
t (response to vj(t)) normalized time t 1z
+ (- cos a n t e-'IT + cos coot) (response to v:(t), first part) Figure 4.11 Resonator phase-step response bu(t);F = Q/(2n), see (4.79).

+ sin w,t e-'/' - sin mot) (response to vy(t), second part).


Hence 1.5

lu
Having defined the detuning frequency Q as wo - w,, (4.79), it holds that sinant = h

sin(w0t - Qt) and consequently that


z' 1.0
i!
0
sin wnt = sin coot cos Qt - cos wot sin Qt . %
2
The output signal (4.88), rewritten in terms of wo and R, is 2 0.5
1
vo(t) = cos wot - K sin wot + K sinmot cos Qt e-'/' -K cos wot sin Qt e-'lT , !?
.3

d,
which simplifies to A
C
A
0.0
a
uo(t) = cos mot(1 - K sin Qt e-'IT) - K sin mot(1 - cos Qt e-'/') . (4.89)

Phase impulse response b ( t ) -0.5


0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Freezing the oscillation wot (i.e. factoring it out), the output signal (4.89) turns into the normalized time t l z
slow-varying phasor
Figure 4.12 Resonator phase-impulse response b(t);F = Q/(2n), see (4.79).
1
V,(t) = --- [l + j ~ ( -1 cos Qte-'/')] , K << 1 .
4
The angle arctan (S{Vo(t))/91{Vo(t)}),normalized to K , is the phase-step response Phase frequency response B(s)
The Laplace transform B(s) = C{b(t)} is found using the Euler formulae
bu(t) = 1 - cos Qt e-"' (step response, Fig. 4.1 1) .
1
Using the property b(t) = (dldt) bu(t), (4.64), we find the impulse response cos at = - (ej"'
2
+ e-jRt) ,
b(t) = (Q sin Rt + Z
e-'I' (impulse response, Fig. 4.12) .
110 Phase noise and feedback theory 4.5 Proof of the beeson formula 111

(a) Resonator tuned at o = on (b) Resonator detuned by !2<< oo

Figure 4.13 Phase-noise transfer function B(s) on the complex plane.

-.-
and the properties 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
normalized frequencyf l fL

Figure 4.14 Modulus squared of the phase-noise transfer function, ~ ~ ( j o(4.76);


)l~,
F = 0/(2n).
L{eat f(t)) = F(s - a ) .

Expanding C {b(t)) , we find


4.5 Proof of the Leeson formula

Owing to the practical need for stabilizing the output amplitude, the oscillator is in-
herently nonlinear. Yet, a linear model is correct for small phase perturbations to the
and finally stationary oscillation. This statement can be justified further if the noisy oscillator is
1
B(s) = -
+ +
s l / t a2t
(transfer function, Fig. 4.13) ,
described by a stochastic equation whose solution converges on the limit cycle. In the
case of a sinusoidal oscillation, the limit cycle is a circle or an ellipse. Of course, a
t ( s + I / t - jLl)(s+l/t+jS2)
smoothly nonlinear system can be linearized for small perturbations around the limit
which is (4.76). cycle. In actual oscillators the phase noise is always a small perturbation.
That said, we will abandon the generic model of Fig. 4.15(a) in favor of the model of
Remark Fig. 4.15(b), which is specific to phase noise. This model describes phase perturbations
The phase-noise bandwidth of the resonator increases when the resonator is detuned around a stationary oscillation, for which AB(jw) = 1 at the oscillation frequency oo.
(Fig. 4.14). This is related to the following facts. The physical quantities of Fig. 4.15(b) are the Laplace transforms of the oscillator's
phase fluctuations. Thus the input signal W(s) models the amplifier's phase noise, or
1. When the resonator is detuned, it holds that (Fig. 4.4) other noise sources referred to the amplifier input. Additionally, W(s) can be used to
introduce resonator fluctuations, or the phase noise of an external signal to which the
oscillator is injection-loclted. In phase space, the amplifier gain is exactly unity because
the amplifier repeats the input phase to the output. The feedback function is B(s),
For a lower slope, the oscillator phase noise is higher. discussed thoroughly in Section 4.4 above.
2. Detuning the resonator, the symmetry of arg P(jw) around the oscillation frequency Describing phase noise, the oscillator transfer function is defined as
is lost. This explains the overshoot seen in Fig. 4.14 for S2 # 0.
3. It may be seen in Fig. 4.1 1 that the step response is faster when the resonator is H(s) = - (definition of H(s)) ,
detuned. *(s)
Phase noise and feedback theory 4.5 Proof of the Leeson formula 113

(a) Voltage space Combining (4.92) and (4.95),


v,($1
out
y( v;
-- - - - - (s1
-a
I
I oscill.
I out

recalling that r = 2 Q/oo = Q/(nvo) = 1loL,(4.18), and using o = 2nf we get

(b) Phase space


and finally

S ) = [I + ( )] s ( 1 (Leeson formula) .

' 1 1
resonator This confirms the Leeson formula (3.19), which was found in Chapter 3 using heuristic
argumentation and physical insight.
Figure 4.15 Derivation of the oscillator phase-noise model (b) from the voltage-noise oscillator
model (a).
4.5.2 * Detuned oscillator
and thus Let us consider the more general case of an oscillator oscillating at the frequency wo =
+ +
on Q, with Q 0. The phase-noise transfer function H(s) is found by substituting
B(s), (4.76), into (4.93). Using the simple property that 1/(1 - N / D ) = D/(D - N ) ,
From the general feedback theory, with the scheme of Fig. 4.15(b) the transfer function we get
is given by
1
H(s) = -. (4.93)
1 - B(s)
Hence the function H(s) is the essence of the Leeson formula (3.19).

Oscillator tuned exactly to the resonator's natural frequency -


(st + 112 + Q2r2
At the exact natural frequency o n , the resonator's phase transfer function is B(s) =
(sr + 1)2 + Q2r2- (ST + 1) - Q 2 t 2 '
l/(sr +l), (4.71). Substituting B(s) into (4.93) we get
and hence
1+ s t
H(s) = ------- (Fig. 4.16(a)) ;
St (sr + + 1)2 Q2r2
(detuned by Q) ,
thus
H(s) = +
s r ( s t 1)

H(S) = (' 'Ir - + 'Ir + jn) (alternative form of (4.101)) ,


+ l/r)
+

sr(s
114 Phase noise and feedback theory 4.5 Proof of the Leeson formula

(a) Resonator tuned at wo (b) Resonator detuned by 52 << w,

Figure 4.1 6 Oscillator phase-noise transfer function H(s) on the complex plane.

and so V '

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0


normalized frequencyf l fL

Figure 4.1 7 Modulus squared of the phase-noise transfer function IH(j w )12, (4.104);
F = S2/(2n).

If the oscillator is pulled by inserting a static phase into the loop, the oscillation
The derivation of 1H(jw)l2 is omitted. frequency is not equal to the resonator's natural frequency. In this condition it holds
Figure 4.16 shows H(s) in the complex plane. When the oscillator is pulled away +
that S2 0. Equation (4.105) indicates that the oscillator phase noise then increases.
from con, the real zero at s = - o ~splits into a pair of complex conjugate zeros at However, if w, is changed by malting the resonator interact with an external reactance
s = - w ~ jS2, leaving a real pole at s = - w ~ in between. The pole at s = 0 in then S2 is unaffected. Thus the phase noise does not increase. The following cases are of
Fig. 4.16 is an ideal integrator in the time domain. This leads to the Leeson ef- interest.
fect, i.e. the divergence of the phase noise in the long run. Seen from the imagi-
nary axis at o >> WL,H(s) appears as a small cluster of two poles and two zeros Quartz oscillator
that null one another, for H(jw) is constant. The resonator is a flywheel that blocks The quartz oscillator is usually pulled by means of an external reactance, as in Fig. 3.7,
the phase fluctuations, since the oscillator is an open loop. Accordingly, the ampli- for S2 is unchanged. Yet the additional loss introduced by the external reactance may
fier phase noise *(s) is repeated as the output. This is the term "1" in the Leeson lower the quality factor and in turn increase the resonator noise bandwidth.
formula (4.100).
The plot of l ~ ( j w ) l shown
~, in Fig. 4.17, reveals that the phase-noise response is a
finction of S2 that increases as S2 increases. This is best shown by the integral Low-noise microwave oscillator
In low-noise schemes such as the Galani oscillator [34], it often happens that the resonator
cannot be tuned, because a phase shifter is used to pull the oscillation frequency. It follows
that S2 # 0 and additional noise will be talten in.

Pound-stabilized oscillator
In the Pound scheme [77], the modulation and detection mechanisms ensure that the
4.5.3 Pulling the oscillator frequency
resonator is tuned at the exact natural frequency onby equating the amplitudes of the
The oscillator can be pulled to a desired frequency in an interval around o n , as shown reflected sidebands. Therefore in this oscillator it holds that S2 = 0 unless a frequency
in subsection 3.1.2. offset is forced intentionally through the control loop.
116 Phase noise and feedback theory

4.6 Frequency-fluctuation spectrum and Allan variance 4.7 kk A different, more general, derivation of the resonator
phase response
If the phase-noise spectrum inside the loop is S+(f ) = bo + b-1 f -' f . . . , the oscil-
lator's fractional frequency spectrum is The derivation of the resonator phase response4 is approached here in a completely
different way, which generalizes the results of Section 4.4 in that:
bo
sJ)(f)=-Zf
Vo
2 b-1
+yf
~o
+-+--.
bo b-ll
4Q2 4 Q 2 f
(4.106)
the resonator equation is left arbitrary;
This is easily seen as follows: it predicts the possibility of a phase-amplitude interaction.

Let us denote by &(t)the resonator's (ordinary) impulse response and by B(jw) its
(derivative) Fourier transform. Further, let us denote by b(t) the resonator's phase-impulse response
and by B(jw) its Fourier transform. We are familiar with the functions P(jo), b(t),
and B(jcl,); b(t) is introduced here for the first time. We will assume that the res-
onator is replaced by a band-pass filter whose frequency response is a sharp peak at
o % wo. In a band-pass filter, the system roots that produce a frequency-selective be-
havior are clustered in two symmetrical regions around fjwo. Additionally, zeros are
present at the origin and at infinity. Such a filter is a generalization of a Q >> 1 res-
onator that can be used to model a variety of physical systems, the most interesting
of which are a resonator having other resonances in the vicinity of the oscillation fre-
The proof extends this result to higher-order terms, in b-2 f -2, etc. quency and a resonator with quasi-degenerate or degenerate resonances at the oscillation
The oscillator's Allan variance is frequency.
Letting

+ +
vi(t) = C O S [ W ~5 ~ qi(t)] = 91 {eJWof
eJr eJa(f)) (input signal) (4.108)
This can be demonstrated by matching (4.106) to the power law S,(f ) = xihifl, and be the input signal, the output signal is then
by identifying the terms; thus
vo(t) = (b * vi)(t) (convolution)

The coefficients ho and h-1 are converted into the Allan variance using Table 1.4 in
Section 1.8. It is worth mentioning that, for reasons detailed in Chapter 2, terms of
higher order than b-2 f -2 cannot be present in the amplifier noise. They can be included
in the formula for the sake of completeness, however, because $(t) models all the phase
fluctuations present in the loop.

Example 4.1. Calculate the Allan variance and deviation of a microwave dielectric Owing to the high signal-to-noise ratio of real oscillators, we can linearize the expression
resonator oscillator (DRO) in which the resonator quality factor is Q = 2500 and the of vo(t) for Iqi(t)l << 1:
+
amplifier noise is S,( f ) = 10-l5 lo-"/ f (white noise -150 d~ rad2/Hz), which
results from F = 4 dB, Po = -20 dB m, and flicker noise - 110 dB rad2/Hz at 1 Hz). &(u)e-JWo"
[1 + jqi(t - ti)] du
Using (4.107),

he analytical method used here was suggested by Charles Greenhall (private communication).
118 Phase noise and feedback theory 4.7 *p11'9$ A different, more general, derivation of the resonator phase response 119

Making the replacement /3(jwo) = POejo,we have The filter functions bc(t) and b,(t) can be rewritten using cos x = f (ejr + e-jx) and
sinx = f (ejx -
52 e-'") :

Using ej" = cos x + j sinx, we observe that Interestingly, (4.1 12) shows the existence of a phase-to-amplitude noise conversion,
which may null in some specific conditions.
The expressions (4.11 1) and (4.1 12) yield naturally the transfer functions
Defining
B(jw) = (PM -+PM conversion) ,
@i(jw)

A(jw) = A,(JO) (PM -+ AM conversion) ,


@i(Jw)
we get where upper case denotes a Fourier transform; B(jw) is the phase transfer function
with which we are familiar, A0(jw) = F{a0(t)) is the normalized output amplitude, and
A(jw) is the phase-to-amplitude conversion function. Combining (4.11 1) and (4.113),
we get

Similarly, combining (4.112) and (4.114), we get


Using 1 + x + j y = (1 + x)eJJ1for small x and y, we get

We aim to identify the fractional amplitude ao(t) and the phase qo(t) at the res-
onator output, malcing an implicit reference to the output signal written as vo(t) = Hence
+
[ l ao(t)]cos[wot + qo(t)], as in (1.6). For this purpose, it is useful to no~~malize
the
input signal so that ejc/30ej0 = 1. This means that we set { = -8 and scale up the
amplitude by a factor 11/30. The output signal thereupon becomes

The above equations (4.1 17) and (4.1 18) apply to the region around w = 0 where phase-
The output phase and amplitude are noise mechanisms take place. The reason comes from the definitions of bc(t) and bs(t)
(see (4.109) and (4.1lo)), which produce up-conversion and down-conversion of b(t),
(PM -+ PM conversion) , respectively, because of the multiplication by a sinusoid. This is illustrated in Fig. 4.18,
where the Fourier transform is replaced by the Laplace transform and shown in the
(PM -+AM conversion) . complex plane. The region of the system function around jwo that contains the roots
generating the band-pass behavior is up-converted to 2jwo and down-converted to dc.
120 Phase noise and feedback theory 4.8 *k Frequency transformations 121

Irk Frequency transformations


The classical theory of filter networks suggests that an arbitrary filter can be obtained
from a low-pass prototype by conformal transformation of the complex variable [43,
Section 14.4, 101, Section 13.4, 106, Section 11.91. The transformation that is most
+
interesting for us is s -+s l l s , because it turns a low-pass filter into a band-pass by
adding a high-pass function obtained from the low-pass function using the transformation
s -+1/s. Here we prefer to show the analogy between band-pass (BP) and a "true" low-
pass (LP), instead of using the dimensionless frequency of the LP prototype.
We start from the single-pole LP function

where the complex variable p is used instead of s to avoid confusion. By making the
replacement

p = - -2 s2 s
+ (LP i BP transformation)

in (4.12I), we find that

Figure 4.18 Frequency-shift band-pass (resonator) to low-pass transformation.

The same thing happens to the symmetrical region, around -j o o . The roots of A(s) and Finally, using r = 2Q/con, thus 2 / =~ w,/Q, we see that (4.123) is equivalent to the
B(s) in the regions around &joo and &2jwo correspond theoretically to phase noise at canonical form (4.25) of the resonator, here repeated:
Fourier frequencies equal to the carrier frequency or twice the carrier frequency, but this
has no practical relevance. Before getting into confusion about this point, one should
recall that A(s) and B(s) refer to the phase behavior of the filter, The phase, though
mathematically defined at any frequency, makes no sense far from the band-pass region It is instructive to learn more about how a frequency transformation maps an LP
because there is virtually no signal. function into a BP function (Fig. 4.19). From (4.122) we get

Input signal tuned exactly to the resonator's natural frequency


When the input signal is tuned exactly to the resonator's natural frequency, coo = on,
it holds that 0 = 0 and hence e i j @ = 1. Owing to the symmetry of P(jw) (subsec-
tion 4.2. I), (4.1 17) and (4.1 18) reduce to Thus, we can deduce the following.

The point p = oo maps into s = 0. This is seen by evaluating the limit of (4.124) for
p -+ 03.
The point p = 0 maps into s = jw,. This is seen in (4.125).
0 +
After expanding p as C jS2, a point p = Z on the real axis maps into
This is the result already found in subsection 4.4.2; the pole pair of P(s) at s = o, & jo,
turns into the pole of B(s) at s = o,.
122 Phase noise and feedback theory 4.8 k k Frequency transformations 123

Figure 4.20 General band-pass (resonator) to low-pass transformation.

thus, using the definition (4.3 1 ) of the dissonance X ,

1 .
P = - JwnX
2
. (4.133)
Figure 4.19 Low-pass to band-pass (resonator) transformation.
Dividing both sides of the above by w~ = w,/(2Q), and recalling that w~ = l / z ,
that is, we get

s = C 41 jw, J1 - X2/w,2 forZ1 < o n ,


which allows us to rewrite the expression (4.121) for the LP equivalent function:
s=C&C
4 1-w,2/C2 for C > w, .
Similarly, a point p = j S2 on the imaginary axis maps into

s=-jfi&jw, \/----
Q2+w2, The natural extension of this is to replace the imaginary variable j x by the cornplex
dissonance
that is ,

s = j S 2 f jw,
J 1 +Q2/w,2 It is to be remarked that the dissonance X , and therefore also the complex dissonance
X, is a warped and normalized frequency and that it is dimensionless. Using X, the
or
equivalent LP filter is

s= jfi=t j~\/1+02/~2.

Replacing s by j w in the transformation (4.122),we get


Finally, for the purposes of understanding the resonator we prefer the inverse transfor-
1 mation, that is, the transformation from band-pass to low-pass. A generalized form of
p = -2 j w , ( ~ - : ) ((4.122)with s = j w ) .
this transformation is shown in Fig. 4.20. The system function has two symmetrical sets
124 Phase noise and feedback theory

of roots generating the band-pass behavior, in the region around jopand in the region
around -j ~ Additionally,
p the band-pass function has one or more zeros at s = 0 and
at s = oo. The frequency transformation moves the &jopregions to the origin, overlap-
5
ping them. The zeros at s = oo are preserved, while the zeros at the origin are moved
to s = oo. The problem with this conformal transformation is that it warps the complex
plane, splitting or overlapping the roots; thus one loses insight.

The basic delay-line oscillator, shown in Fig. 5.1, is an oscillator in which the feedback
path has a delay zd that is independent of frequency. Hence, when the amplifier gain
A = 1, the Barlchausen condition is met at any frequency ol = (2n/td)l with I integer,
that is, a frequency multiple of the free spectral range 2 x 1 ~With. appropriate initial
conditions, stationary oscillation talses place. If the exact condition A = 1 is met in a
frequency range, several oscillation frequencies may coexist.
A real oscillator requires a small-signal gain A > 1 that reduces to unity in appropriate
large-signal conditions. Of course, a selector filter is necessary in order to choose a
mode nz and thus a frequency o,,, = (2x/rd)77z. The filter introduces attenuation at all
frequencies o,with I # nz. For laboratory demonstration purposes, the imperfect gain
flatness as a function of frequency of real amplifiers is sufficient to select a mode,
albeit an arbitrary one. A true band-pass filter is necessary for practical applications.
Using a tunable filter the oscillation frequency can be switched between modes, as in a
synthesizer, in steps A w = 2n/zd.
It will be shown that the laser is a special case of a delay-line oscillator. Addition-
ally, the delay-line oscillator is of great interest for the generation of microwaves and
THz waves from optics because a long delay can be implemented thanks to the high
transparency of some materials, such as silica, CaF2, and MgF2, at 1.55 pm wavelength.
Finally, copper delay-line oscillators are also of technical interest and are comnercially
available for the VHF and UHF bands.
We will analyze the phase-noise mechanism in the delay-line oscillator following the
approach of Chapter 4, which was based on the elementary theory of linear feedback
systems.

Basic delay-line oscillator

The delay-line oscillator can be represented as the feedback model of Fig. 5.1. This is
similar to Fig. 4.6(a) but the resonator is replaced by a delay line of delay zd. The signal
V,(s) allows initial conditions and noise to be introduced into the system, as well as
the driving signal if the oscillator is injection-locked. The noise transfer function, again
defined as
126 Noise in delay-line oscillators and lasers 5 . U a s i c delay-line oscillator 127

model output V, (s)

initial conditions,
noise, or lockmg
,
Vi ($1

I
--
v;(s)
----4s
true
signal I oscillator
I
I
output
I_

....... - .-
-. -. - ............. .- .....-...
4
/
in practice, delay + selector'!
/

/"

. --- - - - --- --- - 1

Figure 5.1 Model of the delay-line oscillator.

is the same as that given in (4.43):


Figure 5.2 Poles of the noise transfer function H(s).The I values, (5.4), are given to the left of
the pole positions.

A delay -cd in the time domain maps into multiplication by e-szd in the Laplace transform.
Thus, in this case, it holds that Interestingly, the derivative

The poles of H(s) are the roots of the denominator, that is, the solutions of 1 - Ae-"'d
+
= 0 with s = cr jw. These poles, which we denote by sl, are shown in Fig. 5.2. They
indicates that a gain change 6A causes the poles to move by 6s, = 6A/(zdA). Thus, the
form an infinite array on a line parallel to the imaginary axis:
poles are shifted horizontally by 60 = 6A/(tdA) if 6A/A is real. From this standpoint,
a fluctuation in the amplifier gain with no associated phase fluctuation does not impact
on the oscillator phase noise. Obviously, the amplitude must be stabilized by a suitable
The poles are in the left-hand half-plane for A < 1, move rightwards as A increases, lie mechanism that does not appear here.
on the imaginary axis for A = 1, and reach the right-hand half-plane for A > 1. The The power and noise are governed by
Barlhausen condition for stationary oscillation is therefore met for A = 1, as expected.
In this condition, the delay-line oscillator can oscillate at any frequency wr = (2n/td)l.
The root diagram of Fig. 5.2 with A = 1 can be regarded as an infinite array of loss-
free resonators, each of which results from a pair of imaginary conjugate poles. When
excited by appropriate initial conditions, each resonator sustains a pure sinusoidal oscil- Using 1 + +
= (1 - A ) ~ 2A and cosx = 1 - 2 sin2(x/2), the above expression
lation of frequency wl = ( 2 n / ~ ~ )asl ,well as dc. At this level of abstraction, there is no becomes
reason to restrict the number of resonators excited simultaneously. Thus, the oscillation
is a linear superposition of dc and sinusoids: 1
(Fig. 5.3) , (5.8)
IH(jw)12 = (1 - A), + 4A c0s(wtd/2)

which is formally similar to the Airy function A(0) = 1/(1 + a sin 0).
This is the Fourier series expansion of an arbitrary periodic waveform. The property of For A = 1 the function I~ ( j w )has / ~a series of singularities (infinities) at w = 2n1/%,
completeness of the Fourier series tells us that the delay-line oscillator can sustain any integer I, which become sharp resonances for A < 1. Therefore, a sub-threshold delay-
periodic waveform. line oscillator can be used as a filter.
Noise in delay-line oscillators and lasers 5.2 Optical resonators 129

mii-ror 1 mirror 2

jT2 I,
>
input round trip output

plane

reference
mii-ror I plane mirror 2

Figure 5.3 Noise transfer fimctio~l( H(jJ'rd)I2,for a delay-line loop with no selection filter.

Optical resonators delay


2,/2

Figure 5.4 Fabry-P6rot etalon.


Fabry--IDbrot interferometer
The Fabry-P6rot (FP) interferometer, also referred to as the FP etalon, is, for reasons not Finally, B(s) = £O(S)/£~(S)is given by
elucidated here, a popular and well-studied optical resonator. The details are available
- T;
in numerous references, among which the author prefers [94,90, 101. Here, we focus on B ( 4 = 1 - ~ 2 ~ - s r 'd
the analogy with the delay-line oscillator. With reference to Fig. 5.4, we will study this
resonator using the Laplace transform of the electric field E(s). Dropping the trivial factor - T ~ ~ - ~ Qand I ~matching
, the double reflection R~ to the
Let us first recall the behavior of a partially transparent mirror. The field £(s) gives gain A, the above is formally identical to (5.3), since it has the same properties.
rise to a transmitted field j TE(s) and a reflected field R£(s). The coefficients R and T The derivation of (5.11) is based on noticing that the output field is the sum of the
+
are nonnegative numbers. By energy conservation, it holds that T2 R2 = 1 and thus field going straight from the input to the output and the fields after 1 , 2 , 3 , . . . , oo round
0 5 T 5 1 and 0 5 R 5 1. The coefficient 'j ' in the transmitted field, corresponding to trips. Thus
a 90" shift, is a consequence of the phase-matching condition at the mirror interface. It
is not relevant for us, since we are interested only in the beam's intensity. An efficient
FP etalon requires that R be close to 1, and consequently that T be close to 0, for from which we find (5.1 1) using the property zr, xi = 1/(1 - x) for lx I < 1.
interference to take place on a large number of round trips, and ultimately for the etalon
to be highly selective in frequency.
Whispering-gallery resonator
Let TI T2 = T2 and Rl R2 = R ~ We . first notice that a round trip introduces a factor
R2e-srd in the electric field. The transfer function is found imagining the circuit to be In this type of resonator, the field is confined in the equatorial region inside a circular
broken at a reference plane and matching the electric field on the two sides: dielectric cavity by total reflection. Whispering-gallery resonance derives from the in-
phase interference of the traveling field with itself after a round trip. For the energy to
be confined, the refraction index must be higher than that of the free space around the
resonator. Although a spherical resonator is the most usual in the traditional literature,
from which we obtain thanks to the energy confinement in the equatorial region the resonator can take various
shapes, e.g. a ring bump on a cylindrical surface or a curved-edge thin disk. Among
lcnown electromagnetic resonators, the optical whispering-gallery resonator exhibits
the highest quality factors. The incredible result Q = 6x 101° at h = 1.55 pm, thus
130 Noise in delay-line oscillators and lasers 5.3 Mode selection 131

-
L
input fiber

resonator section
resonator

Figure 5.6 Amplitude-only selector filter.

The poles of H(s) are the roots of the denominator


output fiber D(s) = 1 - ~ ~ ( s ) e j @ ( ~.) e - ~ ' ~
Figure 5.5 Disk resonator, coupled with prism-shaped optical fibers. Splitting D(s) = 0 into its modulus and argument, we get
Ape-OTd = 1 ,
va Q % 1025,has been reported [40]. The reader should refer to [68, 521 for the physics 8 - W Q = 0 mod 2 n ,
and the applications of this type of resonator.
Figure 5.5 shows an example of a whispering-gallery resonator, coupled by prism- and finally
shaped optical fibers [53]. By inspection of this figure, the matching condition is given by
(5.9) since (5.11) holds in this case also. Radiative and other losses [38] can be included
in the term R ~ This
. is formally sufficient to describe the resonance, even though the
Td
(poles of H(s)) .
physics is quite different.
td Zd

It can be seen that the modulus p affects only the real part a of the pole and thus the
damping; the argument 8 affects only the imaginary part and thus the frequency. The
Mode selection poles of H(s) are located at

A pure delay-line oscillator, with an infinite array of poles on the imaginary axis,
represents a mathematical exercise with scarce practical usefulness. In fact a selector
filter is necessary, which ensures that the Barlthausen condition AP(jw) = 1 is met at Equation (5.19) will be used to analyze some relevant types of selection filter, either of
a privileged frequency. Such a filter ensures that a pair of complex conjugate poles, only theoretical interest or practically realizable.
indicated by subscripts m and -m, lie on the imaginary axis and that all the other
poles sl, I f f m, are located in the left-hand half-plane, R{sl) < 1. Introducing such a
5.3.1 Amplitude-only filter
selector, we replace the feedback function P(s) of (5.2) as follows:
This selector filter is defined as (Fig. 5.6)
for w x f w,,,, thus I = rtm ,
with Bf(s>= (5.20)
c,withO<~<l, elsewhere(1ffm).
Pd(s) = e-s'd (delay line) , (5.14) Such a filter attenuates by a factor E < 1 all signals except those in the vicinity o f f w,,,.
Pf(s) = p(s)eiecS) (selection filter) . (5.15) Needless to say, it is an abstraction that cannot be fully implemented by real-world
devices.
The function Bf(s) has the properties (4.33)-(4.37) seen at the start of subsection 4.2.1. The poles of H(s) are the roots of its denominator, i.e. they are given by
Introducing the selection filter, the transfer function (5.3) becomes

1 - c Ae-s"d = 0 elsewhere (I f fm) .


Noise in delay-line oscillalors and lasers 5.4 The use sf a resonator as a seleetisn filter 133

(a) Amplitude-only filter (b) Phase-only filter By substituting the above Pf(s) into H(s), we get
1
H(s) = 1 _. Aej6sgnwe-std '

The poles of H(s) are the roots of the denominator


D(s) = 1 - AejOsgnoe-s%

By splitting the modulus and the argument of D = 0, we find


Ae-ard = 17

0 sgn w - wzd = 0 mod 2n .


Hence the poles of H(s), shown in Fig. 5.7(b), are located at

In conclusion, the phase-only filter affects only the imaginary part of the pole location,
not the real part. Thus, the phase 0 results in a frequency shift (Fig. 5.7(b)) given by

for w > 0 ,

for w < 0 .

The use of a resonabor as a selection filter


Figure 5.7 Transfer function H(s) of the delay-line oscillator with a selector filter &(s) inserted
in the oscillator loop. (a) Amplitude-only filter. The I values, (5.23), are to the left of the pole Let us provisionally assume that the resonator is tuned at the exact frequency w,,, =
positions. (b) Phase-only filter, (5.26). The value of Aw is $-O/rdat every pole above the real 2n1?2/td of the mode 1 = 172 of interest, which is a permitted frequency of the delay-line
axis and -e/td at every pole below it. oscillator. The feedback hnction is P(s) = Pd(s)Pf(s),with
Pd(s) = e-srd (delay line), (5.28)
Such poles, shown in Fig. 5.7(a) are located at
w17t S
Pf(S) = - (resonator, (4.25) with w, = w,,) . (5.29)
for1 = fnz, thus w = f w , , , Q s2 + ( ~ t/ Q>s
, +
It is important to understand the roles of the delay line and of the filter. The main
technical motivation for the delay-line oscillator is the need for a delay line that exhibits
for 1 # f m , thus w # f w,,, .
high stability. Hence the oscillation frequency must be determined by the delay line,
with at most a minor contribution of the filter's center frequency. For this to be true, it is
Phase-only filter necessary that

The phase-only filter is defined by


because the frequency fluctuations are weighted by the phase slope (dldw) arg B(jo)
of the feedback elements. This is a consequence of the BarIdausen phase condition
where 0 is a constant. This filter shifts the signals by 0 for w > 0 and by -8 for w < 0. arg P ( j o ) = 0. Equation (5.30) is equivalent to
The sign function s g n o is necessary for the condition Pf(s) = P;(s*) to be satisfied.
Once again, this filter is an abstraction.
134 Noise in delay-line oscillators and lasers

where rf is the filter group delay (in other instances denoted by T,), The approximate pole location is found by inserting the above expressions for p(w) and
2Q
Q(w),evaluated at w = wl, into (5.19) with A = 1. The poles are then at sl = 01 jwl,+
tf = - (filter group delay at w = on,) . (5.32) with
Will
Replacing P(s) byPd(s)Pf(s) in H(s) = 1/(1 - A/?@)),under the assumption that
the amplifier gain is A = 1 we find the oscillator transfer function

This function has a pair of complex conjugate zeros at


win
s,, S; EX -- + j w,,,
and a series of complex conjugate poles, to be discussed below.
As a consequence of the condition (5.3 l), the resonator bandwidth w,,, / Q is large
compared with the free spectral range 2n/td. This means that in frequency ranges F hence, they are at
around w,,, and also around -w,,, it holds that I QX 1 << 1 where x is the dissonance; see
(4.3 1). This range F contains a few modes wl around w,,,. In F , the resonator function
Pf(jw) is close to 1. This fact has the following consequences:
1. The poles of H(s) are chiefly determined by the oscillation of arg Pd(s). They are Additionally,we might be interested in expressing the poles as a function of the integer
close to j2nl/td, which is the result already obtained without the selection filter. offset p from the main mode, defined as
2. As an approximation, we can replace Pf(s) by Pf(jw), which can be written in polar 1- m form > 0,thusforl > 0 ,
coordinates as P= (definition of p ) (5.4 1)
1 + 172 for w < 0, thus for 1 < 0 .

The fractional frequency offset is then


with
1
(modulus) ,
and hence
Q(w) = - arctan Qx (phase) ,
w wni
X=--- (dissonance) .
wn,
Figure 5.8 shows the poles of H(s) located on a parabola centered at w,,,. The pole
3. In the vicinity of the oscillation frequency, thus for I(w - w,,)/w,, I << 1/(2 Q) or
frequencies are shifted by -2Qp/(rdm) with respect to the exact mode frequency
equivalently for 1(1 - m)/m 1 << 1/(2 Q) we can approximate the dissonance as
2n 1/td of the delay line. This frequency shift results from the off-resonance phase of
the filter. The negative-frequency part of the complex plane follows by symmetry.
Interestingly, every pair of complex conjugate poles of H(s) can be seen as a separate
This approximation holds only for positive frequencies and thus for positive 1. The resonator (Fig. 4.4 and (4.25) with w,, - wr), whose quality factor Qi derives from
negative-frequency half-plane can be reconstructed using the symmetry r~iles. the filter Q enhanced by positive feedback. The pole pair selected by the resonator,
4. The modulus p(w) and 6 (w) can be hrther approximated as 1 = In, lies on the imaginary axis since the equivalent quality factor is Q,,, = m. Using
Q = - arctan (~{sl}/91{si)), the quality factor associated with the other pole pairs, for
+
which 1 7n, is
- w1,1
Q (w) = -2 Q --------- .
w,,,
Noise in delay-line oscillators and lasers 5.4 The use of a resonator as a selection filter 137

De-tuned resonator
The assumption that the selection filter is tuned to the exact frequency w,,, = 2n m/rd of
the mode 7n of interest is not realistic in practice. Removing this assumption, the filter's
natural frequency

is still close to w,,,. In fact, the frequency error we cannot exceed half the distance
between modes,
-

otherwise the oscillator switches to the contiguous mode. The condition zd >> t f ,
(5.3 I), is still assumed, for the filter function is locally approximated by a linear phase
and a quadratic absolute value. Under these assumptions, we can replace the term
(wl - w,,,)/w,,, of (5.40) with

This has the effect of shifting by we the parabola on which the poles are located and of
introducing a small frequency offset by the addition of

Figure 5.8 Transfer function H ( s ) of the delay-line oscillator with a resonator as the selector to the imaginary part of si. However, this is not a preferred way to pull the oscillator
filter. The frequency shift is Aw = -2Qp/(tdnz). frequency because the detuning may induce mode jump. Instead, in the case of optoelec-
tronic oscillators, it is possible to change the optical delay by using the thermo-refractive
effect, acting on the temperature, or by using the change in dispersion caused by a change
or equivalently in the laser wavelength.

Flat-response band-pass filter


In the domain of telecominunications, complex band-pass filters are widely used; these
exhibit a nearly flat response over the bandwidth (Fig. 5.9), with only a small ripple.
In the presence of noise, these resonance poles become energized by it. Thus, the radio- The experimentalist may be inclined to use one of these filters as the mode selector
frequency spectrum shows a series of sharp lines at frequencies a distance o, from the because they are commercially available, and out of the common belief that flatness is
carrier frequency w,,, , on both sides because ,u takes positive and negative values. These
lines are easily mistaken for competitor oscillation modes, which under our assumptions
they cannot be. As a relevant difference, random excitation causes incoherent oscillation,
and such oscillation is damped. Of course, mode competition and multimode oscillation
can occur, yet under different assumptions on the gain saturation mechanism; this sort
of behavior is commonly found in lasers.
With the operating parameters of actual or conceivable oscillators, the equivalent
quality factor Q1is so high that the noise transfer function ( H ( j o ) ( 2 (5.33),
, appears to
have a series of infinitely narrow frequency impulses S(w). Examples are provided in
scale I 01, \ 0

Section 5.8. Figure 5.9 Low-ripple band-pass filter used in telecormnunications.


Noise in delay-line oscillators and lasers 5.5 Phase-noise response 139

phase-noise
input I I output
1 delay
B(s) = ePzd

-...... .- ........................ .- ...... -. ..................


; in practice, delay + selector \I '\
I
delay selector I, _-A
!

I B(s) =e-STd- B(s) = Il(l + s z f ) I


\
............................................... -. ..................... - i

Figure 5.10 Phase-noise model of the delay-line oscillator.

one of the most desirable frequency attributes for a pass band. Unfortunately, a flat-
response filter is the M ~ O T Schoice
~ for an oscillator. The reason is that the p # 0 poles
Figure 5.11 Phase-noise transfer function H(s) on the co~nplexplane. Here 1 = nz .
are closer to the imaginary axis than they are for a resonator. This has the following
undesirable consequences:
phase step propagates unaffected through the delay line, since it appears at the line output
1. selecting an oscillation frequency is difficult;
after a delay td. Otherwise, we observe that the delay line would have infinite bandwidth,
2. mode jump may occur;
for the group delay is the same at all frequencies and equal to td. Hence the phase, like
3. the close-in noise pealts are maximally high.
any information-related parameter of the signal, taltes a time t d to propagate.
By substituting B(s) = ePzdin H(s) = 1/(1 - B(s)), we get
Phase-noise response
1
H(s) = 1 - e-szd (delay-line oscillator) .
Thanlts to a gain control mechanism and to the appropriate initial conditions, a delay-time
oscillator oscillates steadily at the frequency w,,, = 2nm/td. We are now able to analyze The function H(s) has an infinite number ofpoles, which are the roots of the denominator.
the phase-noise transfer function, already defined in Chapter 4 as H(s) = cD(s)/!J!(s), Recalling from (5.41) that the integer mode offset is given by (5.41),
(4.9 1). The phase-noise model of such an oscillator is shown in Fig. 5.10. In this circuit,
all the signals are Laplace transforms of the phase fluctuations present in the oscillator.
The input signal !J!(s) is the phase noise of the oscillator loop components, referred to .={ 1 - 1.12
1 + rn
for w > 0, thus for 1 > 0 ,
for w < 0, thus for 2 < 0 ,
the amplifier input. Hence the amplifier is rendered free from noise. The amplifier gain
is equal to unity because the amplifier just "copies" the input phase to the output. Thus we find that the poles (Fig. 5.11) are located at
it holds that
2n
1 s, = j - p for integer p . (5.5 1)
H(s) = - td
1 - B(s) '
which is the same as (4.93), obtained in the same way. In the spectral domain, the square modulus of the phase-noise transfer function is
1
(Fig. 5.12) . (5.52)
No selection filter
In this case, the phase feedback function is Interestingly, 1H(jw)l2 has a series of minima IH(jw)liin = at w = 2n/rd, a
w = 2x/rd, w = 2n/td, etc., i.e. at odd multiples of half the free spectral range.
At these minima, the oscillator phase noise is 6 dB lower than the loop internal noise
which represents a delay line independent of the carrier frequency. A proof of (5.49) can (thermal, shot, etc.) referred to the carrier power. The physical meaning of these low
be obtained with the the phase-step method introduced in Section 4.4. In this case, the minima is that the noise is sunk by the neighboring poles.
Noise in delay-line oscillators and lasers 5.5 Phase-noise response

We first focus on the low-pass phase filter Bf (s). Close to the imaginary axis, we can
approximate Bf (s) as Bf (jo), which is written in polar form as

with

Using tf = 2Q/w,, and otf << 1, we malce the approximations

Figure 5.12 Phase-noise transfer function l H ( jf -cd)12 for a delay-line oscillator with no selection
filter.

A resonator as the selection filter


As the equation 23 = 0 is equivalent to 1 - B(s) = 0, the poles of H(s) are the solutions
Let the delay-line oscillator oscillate at the frequency w,, = 2nm/td, owing to a res- of
onator in the feedback loop tuned to the exact frequency w,,,. We will analyze the
phase-noise transfer function H(s) = \Zl(s)/@(s) under this condition. In the loop, it
holds that H(s) = 1/(1 - B(s)), (4.93). Introducing a resonator as the selector filter, the
feedback function is which splits into modulus and argument:

p(w)e-atd = 1 (modulus)
Q(w) - wtd = 0 mod 2n (argument) .
with
The modulus condition yields
Bd(s) = eWssd (delay) ,
1
Bf(s) = - (filter) .
1 stf+
Expanding H(s), we find that

The function H(s) has a real zero at s = - 1/tf = --wit,/(2 Q), a pole at s = 0, and
2~~ -
o2 o2
a series of complex conjugate poles in the left-hand half-plane, discussed below. By 0 = -- for 2 ~ ' - << 1 .
+
definition, the poles are the roots of the denominator 23 = 1 s t f - e-s'd of (5.56). -G oil 4 1

It was shown earlier that, for technical reasons, it is necessary that t d >> tf = 2 Q/wnl. The phase condition yields
As a consequence, the poles are expected to be close to the position already found in
o
the absence of the selector, that is, close to s = j 2 n p / t d . In other words, the selector mr)td+2Q- =0 mod 2 n ,
has only a small effect on B(s) in a region around the origin that contains just a few pole ~ 1 1 2

pairs. This is exactly the same situation that we found when we were searching for the 2n 2 Q o
w=-p--- forintegerp.
poles of H(s) transposed from fw,,, to the origin. td td on2
142 Noise in delay-line oscillators and lasers 5.6 Phase noise i n lasers 143

frequency, Hz
: : : : I : : : : I : : : I
Figure 5.13 Phase-noise transfer function H(s) of the delay-line oscillator with a resonator as the 103 10" lo5 lo6
= (2n/td)nz and o = -(2 ~ ~ / t d )n( ~~) /~ .
selector filter. Here o,,,
Figure 5.14 Phase transfer filnction I H (f)12
~ for a delay-line oscillator with selector, evaluated
for td= 20 ps, m = 2 x 1O5 (thus v,,, = 10 GHz), and Q = 2000 (the data of Example 5.1 in
Combining (5.63) and (5.64), we find that the poles are located at Section 5.8).

mirror 1 mirror 2

Additionally, replacing w,/w,,, by ~ we Z


obtain
jrl : gain mediuln
-
Z[y);
/ M &o
&i
>
input output
I
initial conditions
or noise
I .
Figure 5.13 shows the phase-noise transfer function H(s) on the complex plane. The refeience
plane
pole at the origin represents a pure integrator in the time domain and causes the Leeson
effect. The other poles are on a horizontal parabola, centered on the origin, at small
negative distances o a -p2 from the imaginary axis. It is easily seen by comparing - - -mirror 1
- - - - - - - .. - - -mirror 2
- - - - - - - -.
, C

Fig. 5.13 with Fig. 4.4 that each complex conjugate pair is equivalent to a resonator of
quality factor '
input / I
I
delay
-9 I
I

which can be rewritten as

Figure 5.15 A laser can be regarded as a feedback system.

because we have w,, = 2n,u/rd.


i.e. odd multiples of half the delay. There the oscillator phase noise is 6 dB lower than
From (5.56), after some lengthy manipulations one obtains
the loop internal noise (thermal, shot, etc.) referred to the carrier power. At these low
IH(jw)l2 =
1 + w2t,2 (Fig. 5.14) . (5.69)
minima, the noise is swamped by the neighboring poles.
2 - 2coswy + w2r: + 2wrfsinwrd
The function 113(jw)12 shows a series of sharp peaks at w = 2 n ~ l - ror~ f = p/rd, Phase noise in lasers
integer / L .The phase noise talcen in is enhanced in the vicinity of these frequencies. The
pealcs derive from the poles of H(s) close to the imaginary axis yet not on the axis, for In Fig. 5.15 the optical system of a laser and the equivalent electric circuit are shown. This
they are not the signature of competing oscillation modes. The function ~H(jw)l~ has figure refers to the classical two-mirror laser, that is, a Fabry-Pkrot (FP) interferometer
a series of minima 1H(jw)lii, = $ at w = 2n/rd, o = 2 i ~ l - w
1 3
r ~=~ $ 2n/rd, etc., with a gain medium inserted between the two mirrors. Gas lasers and semiconductor
144 Noise Cn delay-line oscillators and lasers 5.7 Close-in noise spectra and Allan variance 145

lasers are of this type. Another structure is common in which the cavity is a ring, for of Fig. 5.16(a), o2, . . . , ws are in the mode competition because they lie in the region
example a whispering-gallery resonator or an optical-fiber loop. The arguments given in where the small-signal loop gain is higher than unity, whereas ol does not. It can be
subsection 5.2.2 should have made clear that for our purposes the ring structure is fully seen that w3 wins.
equivalent to the FP structure. Multimode oscillation requires that the Barhausen condition AP(jw) = 1 is met at
The transfer function H(s) = Eo(s)/Ei(s) is found immediately by comparing Fig. 5.15 more than one frequency. This is found in some lasers where the gain bandwidth is
(the laser) with Fig. 5.4 (the Fabry-Perot etalon). Therefore, letting R1 R2 = R ~ , due to the inhomogeneous broadening mechanism. In this case, the gain is provided
by a cluster of energy levels, each of which is narrow enough for the power to sink a
population selectively (Fig. 5.16(b)). When oscillation takes place at a given frequency,
the active population decreases only near this frequency. Thus oscillation can also arise
The Barlchausen condition is met at the frequencies cq = 2nl/Td if A R = ~ 1, thus if the at the other frequencies where the phase condition is met and the small-signal gain is
gain is sufficientto compensate for the reflection loss at the mirror. greater than unity. In the example of Fig. 5.16(b), oscillation takes place at w2, o3, q,
As amplification derives from the stimulated-emission mechanism in pumped atoms, and a s , yet not at wl because the small-signal gain is insufficient.
the gain cannot be constant with frequency. In virtually all lasers, it turns out that the
bandwidth of the laser gain is greater than the free spectral range of the optical cavity.
For our analysis, this will be fixed by taking a constant gain A and pure delay e-S'd and 5.7 Close-in noise spectra and Allan variance
by moving the gain dependence on frequency to a function Bf(s):
From the definition of H(s), the oscillator phase-noise spectrum is
S,(f f S+(f ,
= l ~ ( j)12 (5.72)
This is the case analyzed in Section 5.4, where the mode-selector filter was introduced. where S+(f ) is the phase noise of the amplifier. A power law is generally not suitable
In conclusion, the laser is equivalent to the delay-line oscillator of Fig. 5.1, and thus for modeling the phase-noise spectrum of the delay-line oscillator because of the non-
the framework of Sections 5.3-5.5 and 5.7 applies to it. polynomial nature of H(s). A power law can only be used at low frequencies, where
IH(jf ) l2 is approximated as
5.6.1 Single-mode and multimode oscillation
In a single-mode laser (Fig. 5.16(a)), when the signal exceeds a threshold the gain
drops more or less uniformly, and smoothly, in the pass band. This is found in quantum still under the assumption that Q >> ~ fHence,
. the oscillator phase noise is
amplifiers, i.e. masers and lasers. Oscillation sinks atoms from the pumped level and
thus the active population is reduced. With these amplifiers the gain condition can be met
at only one frequency. Among the modes at which the loop phase is zero, the oscillation
frequency is generally that at whichthe small-signal gain is the highest. In the example

(a) Single-mode laser (b) Multimode laser


Introducing the amplifier power-law model S+(f) = bo + b- /f , which is limited to
white and flicker noise, we get
arg AP = 0
ajg@t,g+in large signal

The fractional-frequency fluctuation spectrum S,,(f ) is found by using S,,(f ) =


( f 2/v,:,)s,(f 1. Thus
I :
,
:
,
:
.
:
.
:
,

\ IAPI=l
arg AP = 0
After matching (5.77) to the power law S,(f) = Cih if", we find

Figure 5.16 Different types of gain saturation in lasers.


Noise in delay-line oscillators and lasers 5.8 Examples 149

Table 5.2 Resonance parameter's for the SAW oscillator of Example 5.2 Focusing our attention on the white and flicker frequency noise, the oscillator phase
noise is
2.5x10-~
S,(f) =
f2
+8x1W5f3

The spectrum of the fractional-frequency fluctuation is

and the Allan variance is

1o -~
thus
10-~
1.4 x 10-l3
oJ,( t ) = (white frequency noise)
10-lo fi
o;,(r) = 1.3 x 1 0-l2 (flicker frequency noise) .
10-12

10-14

10-16
10 102 1 o3 lo4 1o5 106
Figure 5.19 Expected phase-noise spectrum S,( f ) for a SAW delay-line oscillator with selector.
The parameters are t d = 5 ps, nz = 4500 (thus v,,,= 900 MHz), and
Q = 80; bo = -145 rad2/Hz, bPl = -130 db rad2/Hz. The data refer to Example 5.2.

Example5.2. SAW delay-line oscillator. We will analyze a surface acoustic wave (SAW)
delay-line oscillator using the following parameters:

td = 5 ps 15 rnm SAW, sound speed 3 l a d s


l / t d = 200 kHz mode spacing
v,,, = 900 MHz oscillation frequency; v,,, = m Itd
TYZ = 4500 mode order
Q = 80 selector filter (LC filter)
bo = 3.2 x 10-15 white phase noise, - 145 r a d 2 / ~ z
b F l = 10-l3 z f = 1 Hz.
flicker phase noise, - 130 dB r a d 2 / ~ at

These values relate to GSM mobile phones, which operate in the 900 MHz band with a
channel spacing of 200 kHz. By tuning the selection filter, this oscillator can replace a
synthesizer that worlcs at the appropriate frequency, with an appropriate 200 1sHz step.
-'
The amplifier corner frequency, at which b-l f = bo, is f,= 3 1.6 Hz. Figure 5.19
shows the expected phase noise S,(f). The neighboring resonances show up at
f = l / t d = 200 kHz and multiples thereof. Table 5.2 shows the relevant resonance
parameters.
6.1 General guidelines

be driven at very low power, say 10-20 pW, for the best long-term stability. The quality
factor of a dielectric resonator can be 1000 or more, depending on its size and frequency.
And so on for other resonator types and for the amplifiers. Oscillators similar to those
encountered in the past may have a similar spectrum or be surprisingly different.

Parametric estimation of the spectrum


Spectra are in most cases shown as Y (f), which is to be converted into S,( f ) using
2( f ) = f S,( f ). Then, we have to match the phase-noise spectrum to the polynomial

A more serious title for this chapter could have been The reverse engineering ofphase
noise. When a computer hacker gets into trouble, he or she tacltles the root of the problem,
by reading the source code, patching the kernel, and recompiling the system rather than
just rebooting the machine. There are at least two serious reasons for choosing such a in order to identify the coefficients bi. A term bi f i on a log-log plot appears as a straight
radical approach. The first reason is to acquire a deeper understanding of the system line; see the following table.
being investigated, i.e., to "solve the puzzle," and the self-confidencethat goes along with
this. Such an understanding ultimately results in an improvement in the performance
Noise type Term Slope
and the reliability of the entire system. The second reason is that a radical approach
is ultimately the sinzplest and perhaps the only legal way to solve the problem. So white phase bo 0 0
one should solve, rather than find a provisional remedy. This book is far less ambitious.
Nonetheless, having been frustrated for years by the insufficienttechnical documentation
flicker phase b-l f -' -1 - 10 dB/dec

white frequency b-2 f -2 -2 -20 dB/dec


of oscillators, one begins to realize that there is a lot of information written between the flicker frequency b-3 f -3 -3 -30 dB/dec
lines in invisible ink. Yet it is not that invisible! By hacking the technical information, frequency random walk b-4 f -4 -4 -40 dB/dec
we can deduce - or at least guess - the relevant parameters, such as Po, Q, fL, the
amplifier's l/f noise, etc. The need to guess the internal technology is a source of
difficulty, inconsistency, and frustration. How to cope with this is part of the message Terms with negative slope higher than -4 can be present. The actual spectra are of the
addressed to the reader. form
The major problem with reverse engineering is that it is more of an art than a science
and relies heavily on a person's experience. Reverse engineering is not the kind of
thing that can be reduced to a single comprehensive procedure. The examples provided
illustrate how the author tacltles the problem. Yet, even the choice of example reflects where the terms sj( f ) account for mains residuals (50 Hz or 60 Hz and multiples),
an author's experience. for bumps due to feedback, and for other stray phenomena. Numerous figures in this
A related problem is that reverse engineering is not free from errors. The reader chapter provide examples of actual phase-noise spectra. The mathematical process of
should be aware that this general statement, unfortunately yet inevitably, also applies to matching the spectrum to a model is calledparametric estimation [75,56]. Some a priori
this chapter, for which the author apologizes in advance. knowledge of the nature of the stray signals may be necessary to match the complete
+
model ~ y = bi- f~ Cjsj( f ) to the observed spectrum. Although (almost) only the
power-law coefficients bi are relevant to our analysis, the terms sj( f ) are essential in
6.I General guidelines that they reduce the bias and the residuals of the estimation.
Whereas computers provide accuracy, a general parametric estimator is not trivial
6.1 .I to implement. Conversely, even with little training the human eye is very efficient in
Inspection on the data sheet
extracting the useful information from a spectrum, filtering out the stray signals, and
The first step consists of reading carefully the data sheet, focusing on the resonator and getting a good approximation. The inspection of a log-logplot by sliding a (preferably
on the amplifier technology, and assembling related facts in one's mind. For example, transparent) old-faslzioned set square against a ruler (Fig. 6.1) before using a computer
a 5 MHz quartz oscillator may have a quality factor in excess of lo6, but then it must is strongly recommended. The author has been seen numerous times in the exhibit areas
152 Qscillalsr hacking 6.1 General guidelines 153

from averaging the two worked-out spectra. Physical judgment should be used to assign
unequal weights to the two analyses.

6.1.3 Interpretation
This part of the process starts from identification of the spectrum type among those
previously analyzed. After that we need to learn about the specific oscillator, which is a
unique case. Only a few rules are given here. The reader should refer to the examples
given in this chapter then analyze his or her own spectra.
As a general rule, one should analyze the spectrum from the right-hand side to the
left, thus from white phase noise to frequency flicker or to random walk.
Starting from the white phase noise region we evaluate the power Po at the input of
the sustaining amplifier, using S,( f ) = bo = FkTo/Po, (2.30). Thus

ourier frequency, Hz

v
One may admit a noise figure F = 1.25 (1 dB) for conventional amplifiers, and F = 3.2
(5 dB) for noise-corrected amplifiers owing to the input power splitter. Thanks to the
Figure 6.1 Inspection on a log-log plot by sliding a set square and ruler provides a quick and gain of the sustaining amplifier, the white noise of the output buffer can generally be
surprisingly accurate parametric estimation of the spectrum. The spectrum is that of Fig. 6.9,
neglected.
used with the permission of the IEEE, 2007.
The next step is to evaluate fc (the flicker frequency of the sustaining amplifier) and
fL, in order of occurrence from right to left. It is then necessary to guess the oscillator
of international conferences in the process of sliding credit cards on a phase-noise sub-type (Figs. 3.12 and 3.13). A major difficulty is that of understanding whether the
spectrum and drawing a straight-line approximation with a sharp pencil. Even this poor oscillator stability derives from the Leeson effect or from resonator fluctuation. Inverting
replacement for proper drawing tools has proved to be surprisingly useful. (3.2 I), the Leeson frequency gives a quality factor
The best way to analyze the spectrum is to work with exact slopes f ', l / f , l / f ', etc.
obtained from the coordinate frame using the longest possible baseline. The reference
slope is compared with the curve by sliding the set square until its side coincides with
The corner frequency fc reveals the phase flickering of the amplifier:
the portion of the spectrum having the same slope. Generally, at the corner between two
straight lines the true spectrum is found to be 3 dB above the corner point. This is due
to one of the following reasons.
In some weird cases the 1/f noise of the output buffer shows up. This has been observed
in ultra-stable 5 MHz quartz oscillators and in noise-degeneration microwave oscillators.
1. The difference in slope is 1, i.e., 10 dBIdec, at the corner point. When this occurs,
there are two independent random processes whose spectra take the same value
Allan variance
(bi f' = bi+l f ' + l ) at the corner point. The Allan variance of the fractional frequency fluctuation (Table 1.4), i.e., the oscillator
2. The difference in slope is 2, i.e., 20 dB1dec. When this occurs, at the corner point
stability, is related to the phase noise by
we have bi f = bi+' This implies that a single noise process is filtered, owing
to the Leeson effect; the factor 2 (3 dB) at the corner point results from a single real
zero of the complex transfer function.
Comparing the Allan variance calculated in this way with the Allan variance reported on
In some cases the difference between the spectrum and the straight-line approximation the data sheet or with the measured Allan variance is often a source of frustration, for a
differs from 3 dB at the corner point. Then two estimations are needed, the first based number of reasons. First, the variance is always approximate because the integral (1.8 1)
on the straight-line fitting and the second based on the 3 dB difference between the (or (1 39) for the modified Allan variance) may not converge without the introduction
straight lines and the true spectrum at the corner frequencies. The best estimate results of a low-pass filter. Of course, the bandwidth and roll-off law of this filter will affect
the result. Second, and more generally, the measurement results are sensitive to the
experimental method and to the operating parameter. Finally, time-domain measurements
and spectrum measurements are considered as two separate worlds. This occurs because
the two domains tend not to involve the same people and because very few oscillators
are measured careklly in both domains, overlapping the time frame. Of course, having Leeson effect?
consistent results involves an effort that is seldom made. In conclusion, the reader should
not be discouraged by any lack of consistency between spectra and variances.

6.2 evaluate
About the examples of phase-noise spectra

The phase-noise spectra used in this chapter, with very few exceptions, were taken from
a manufacturer's website or data sheet, or from journal articles. When it was necessary
to redraw a spectrum, the original was imitated as far as possible. This choice has the
advantage of training the reader to the real world, at the inevitable cost of making direct
comparisons that are sometimes uncomfortable.
All the formulae of this book refer to S,( f), while the manufacturers prefer Z (f).
Therefore, the vertical scales of most spectra have been overwritten in order to facilitate
1
the reader. Recalling that by definition Z ( f ) = S,( f ), the conversion from the tech-
nical unit dBc/Hz to the §I unit dB r a d 2 / ~ requires
z the addition of 3 dB. This explains
%&/
why the numbers on the vertical scale end in "7." $5

All the figures carry the author's deductions. Most data fitting was done graphically, khe Leeson effect7
as in Fig. 6.1, working on an enlarged copy of the original plot and interpolating sharp
thin lines with patience. The computer version of the figure was prepared only at the
end of the process. In at least one case (Fig. 6.16), however, the spectrum was worked
out directly on a computer, yet enlarging the picture to full size on a 23 inch screen.
The interpretations proposed in this chapter are fully independent of the manufacturer's
Figure 6.2 Interpretation of the phase noise in ultra-stable quartz oscillators. Adapted from [83]
feedback, when provided. When a piece of information comes from the manufacturer, and used with the permission of the IEEE, 2007.
which occurs in a very few cases, this is clearly stated.
In ultra-stable 5-10 MHz oscillators, phase flickering (b-l /f ) is always present in the
phase-noise spectrum. This fact is discussed below, with the help of Fig. 6.2.
Understanding the quartz oscillator
1. The phase-noise plot changes slope by 2 (from I /f to 1/f ) at f = . We want toft
The piezoelectric quartz is certainly the oscillator type that has the largest production, lcnow whether this corner point is the Leeson frequency or something else.
the largest variety of applications, and the highest economic impact; thus it takes the first 2. Owing to the high Q value and the low carrier frequency vo, and thus the low
place in our analysis. We will focus only on ultra-stable quartz oscillators. This choice value of fL, the oscillator loop is expected to produce an observable I / f phase
is like testing the technology of future production cars on formula-one racing models. noise.
Most of what we can learn is found there. 3. The oscillator loop requires high isolation from the output to prevent the load from
A common signature is found in a number of phase-noise spectra and this deserves affecting the oscillation frequency, so a buffer is necessary. Design practice suggests
to be introduced before the case studies [83]. The procedure detailed in this section that there can be three buffer stages.
should help the reader to understand the interplay between a phase-noise spectrum and 4. The l / f phase noise of cascaded amplifiers adds up directly; the contributions
the oscillator parameters; it leads to the conclusion that the amplifier's phase noise are not divided by the gain of the preceding stages. This is explained in detail in
transformed into frequency noise via the Leeson effect is less than the fluctuation in the subsection 2.3.2. Thus, the total flicker b-l /f is the sum of the sustaining-amplifier
resonator's natural frequency. noise plus the buffer noise.
156 Oscillator hackin

5. With three buffers whose l / f noise is similar to that of the sustaining amplifier,
the noise of the sustaining amplifier is about one-quarter of the total flicker, that
is, 6 dB lower. The difference could be even greater than 6 dB, because the output
amplifier can be more complex and because superior technology would be used in
the sustaining amplifier rather than in the buffers.
6. The Leeson effect takes place inside the oscillator loop, before buffering. After
taking away the buffer noise we observe that the sustaining-amplifier l / f noise
crosses the 1/f 3 noise at f = f[. Once again, we want to laow whether this corner
point is the Leeson frequency or something else.
7. Iff[ is indeed the Leeson frequency then we are able to calculate the quality factor
using Q = v0/(2fL). We denote the result by Qs, where the subscript "s" stands for
spectrum,
8. Leaving the noise spectrum aside for a while, we estimate the quality factor on the
grounds of the technology involved. We denote this value by Qt, where the subscript
"t" stands for technology.
9. In all cases analyzed, it turns out that Qt << Q, and that the ratio Qt/Qs is far
too high for Qs to be the resonator's quality factor, even accounting for the large
102 :
I
103 104
f: = 250 Hz
Fourier frequency, Hz
uncertainty in guessing Q,. Thus, we have to trust the technology and admit that the
Leeson frequency is fL = vo/(2Qt) < f[. Figure 6.3 Phase noise of the oven-controlled 5 MHz quartz oscillator Oscilloquartz OCXO
10. The final conclusion is that resonator instability overrides the Leeson effect. This 8600, with a preliminary (wrong) interpretation. Courtesy of Oscilloquartz SA, 2007. The
conclusion, of course, comes only after a thorough analysis of the Leeson effect. interpretation, comments, and mistakes are those of the present author.

The process described relies on the ability to estimate the resonator's quality factor. polynomial c:=-,
bi f', with
Experience indicates that the product voQ is a technical constant for piezoelectric quartz bo -155 dB 3.2 x 10-l6 rad2/Hz
resonators, in the range from 1 x 10" to 2 x 1013.As a matter of fact, the highest values b-1 -131 dB 7.9 x lo-14 rad2/Hz
are found in 5 MHz resonators. When the resonator is loaded by the rest of the circuit, b-2 (not visible)
its Q value is reduced by a factor 0.5-0.8. The actual value depends essentially on the b-3 -124 dB 4 10-l3 rad2/HZ
sustaining amplifier and, in turn, on the frequency and on the available technology. A
lot of data are available from [36, 62, 1041, and from our early attempts to measure A preliminary interpretation of the spectrum is the following.
the resonator frequency stability [84]. The oscillators we have selected for analysis 1. The white phase noise is bo = 3.2 x 10-l6 r a d 2 / ~ (-
z 155 dB rad2/I3z). Thus
exhibit the highest available stability and are primarily intended for high-end scientific
applications. This background encourages one to trust the published data.

assuming that F = 1 dB. This power level is consistent with one's general experience
of 5-10 MHz oscillators, in which the power is kept low for best stability.
Quartz oscillators 2. The value f; 2i 2.2 Hz, if it corresponds to the Leeson frequency, gives a quality
factor

Oscilloquartz OCXO 8600 (5 MHz AT-cut BVA)


Figure 6.3 shows the phase-noise spectrum of the oven-controlled quartz oscillator The question is whether f; is in fact the Leeson frequency.
Oscilloquartz OCXO 8600. This oscillator has been chosen as an example because of Item 1 above is correct because the effect of the output buffer on the white noise
its outstanding stability in the 0.1-10 s region. The spectrum is of type 2 (fL < f,),as is divided by the gain of the sustaining amplifier, thus we believe that it is negligible.
is typical of low-frequency oscillators with high-Q resonators. The plot is fitted by the The white noise originates in the series resistance of the resonator and in the sustaining
Oscillator hacking 6.4 Quafiz oscillators 159

5. There is no doubt that Q > 5.6 x 105. This means that the Leeson effect is hidden
and that we are not able to calculate Q. Henceforth, our laowledge about Q relies
only upon experience and on the literature.
Oscilloquartz OCXO 8600 6. Let us guess that Q = 2.5 x 106, reduced to Q = 1.8 x 1o6 (loaded) by the dissipa-
resonator instability
b-, = -1 24 dB rad 2 / ~ z
5 MHz OCXO tive effect of the amplifier input. It follows that
1

This indicates that the oscillator frequency fliclter b-3 2: 3.2 x 1013 is due to fre-
quency flickering in the resonator. The Leeson effect is hidden by the resonator
instability and by the phase noise of the output buffer. Thus the corner frequency
f{ 2: 4.5 Hz of Fig. 6.4 is not the Leeson frequency.
7. The fliclter noise of the sustaining amplifier combined with the (guessed) Leeson
frequency gives the stability limit of the electronics of the oscillator. This is the solid
: ; 10 \ :' I 102 103 104 line f -3 in Fig. 6.4, some 10 dB below the phase noise.
fL= 1.4 Hz f;= 4.5 HZ. I Fourier frequency, Hz 8. The frequency fliclter expressed as the Allan variance (Table 1.4) is
(guessed) fc = 63 Hz

Figure 6.4 Phase noise of the oven-controlled quartz oscillator Oscilloquartz OCXO 8600, with
revised interpretation. Courtesy of Oscilloquartz SA, 2007. The interpretation, comments and
any rnistaltes are those of the present author.

thus
amplifier input. Conversely, the fliclcer noise of the output buffer adds according to the
rules given in Section 2.4; see (2.5 1). Thus, we change the interpretation as follows. The
spectrum with the author's comments is shown in Fig. 6.4. o;(r) -. 2.2 x lo-" and so o,.(r) 2 1.5 x 10-l3 .

1. Let us assume that the l / f noise of the sustaining amplifier is one-quarter of the
total l / f noise, guessing that there are three stages between the sustaining amplifier This is lower than the value oy(r) < 3 x 10-l3 for 0.2 s 5 r 5 30 s given in the
and the output and that the l / f noise of each stage is at least equal to that of the specifications. As a general rule, the actual Allan variance is higher than the value
sustaining amplifier. calculated in this way because it results from an integral that taltes in all the noise
2. Talting away 6 dB, the new estimate of the sustaining-amplifier fliclcer is phenomena, not just hwl/f. Yet, this incompleteness of the integral is not suf-
ficient to explain the discrepancy. Of course, we should also bear in mind that
specifications are intended to be conservative; a sample can be better than the
specifications.
Accordingly, the new estimated values are f,2: 63 Hz for the corner frequency of
the sustaining amplifier and
Qscilloquap%z 8607 (5 MHz SG-cut BVA)
This oscillator is similar to the 8600 considered above, yet the short-term stability is
for the Leeson frequency. The latter is still provisional. further improved. Figure 6.5 shows the phase-noise spectrum, together with the spec-
-.
3. Using f[ = 4.5 Hz in fL= vo/(2Q), we get Qs 5.6 x lo5. ifications (bold upright crosses). The spectrum is again of type 2 (fL < f,), typical
4. Qs z 5.6 x lo5 is far too low a value for a top-technology 5 MHz oscillator. The of low-frequency oscillators with high-Q resonators. The f-3 noise is hardly vis-
literature suggests that the product vo Q is between 1013 and 2 x 1013. ible on the left-hand side of the S,(f) plot, which starts from 1 Hz. Thus curve
160 OsciPPlalor hacking 6.4 Quartz oscillators 161

The results are


-67 -153 dB 5.0 x 10-l6 rad2/Hz
Oscilloquartz 8607 bo
b- 1 -132.5 dB 5.6 x 10-l4 rad2/~z
5 MHz OCXO
b-2 (not visible)
b-3 -128.5 dB 1.4~10-l3 rad2/~z

We interpret the spectrum as follows.

1. The white phase noise is bo = 5 x 10-l6 r a d 2 / ~ zThus


.

assuming that F = 1 dB. As expected in this type of oscillator, the power is kept low
for best stability.
2. The l / f noise shows up in almost two decades. The Leeson effect (the slope changes
by 2, f 0 -+ l/f2 or l/f -+ l/f3) occurs at a lower frequency. In the l/f region
1 10 I@ 103 104 16 106
we observe the noise of the sustaining amplifier and buffer, which adds up (subsec-
Fourier frequency, Hz
tion 3.3.2).
Figure 6.5 Phase noise of the oven-controlled quartz oscillator Oscilloquartz OCXO 8607. 3. Assuming that there are four cascaded amplifiers, i.e, a sustaining amplifier and three
Courtesy of Oscilloquartz SA, 2007. The interpretation, comments, and any mistakes are those buffers, the contribution of the sustaining amplifier is one-quarter of the total. Taking
of the present author. away 6 dB, we find

fitting is impractical, and we have to rely on the specification table, which is given
below. thus f, = 3.2 Hz.
4. The 1/f line crosses the flicker at f = f{ = 3.2 Hz, which is suspected to be the
Leeson frequency.
f , Hz S,, dB r a d 2 / ~ z 5. Substituting f{ = 3.2 Hz into Q = vo/(2fL), we find Qs = 7.9 x lo5.
6. The value Qs = 7.9 x lo5 is far too low for it to be the resonator quality factor. The
technology suggests Qt = 2 x lo6.
7. Using Qt = 2 x lo6 as the quality factor in fL = vo/(2Q), we find fL = 1.25 Hz.
This indicates that the oscillator frequency flicker (b-31 f in the phase-noise spec-
trum) is not due to the Leeson effect. Rather, it is due to the l / f fluctuation of the
resonator's natural frequency.
Loolting at the spectrum and at the Allan variance, it is clear that at f = 1 Hz and 8. The frequency flicker expressed as the Allan variance (Table 1.4), is
f = 10 Hz the terms b-3 f -3 and b- f -l determine S, (f ), with at most a minor
contribution of bo. It is also clear that S,( f )I l i d r 2 bO and that there is no visible
b-2 f -2. Thus b-3 and b- are obtained by solving
thus
(b-3 f -'4- b-if -I) = 0-12.7
(1 Hz spec.),
o;(r) E 7.83 x and so ~ ~ , (Er 8.85
) x 10-l4 .
- 10-14.2
(b-3 f - 3 +b-lf-l +bO)lOHz- (10 Hz spec.) ,
This is close to the specification for the lowest-noise option, given as a;,(r) = 1 x
bo = 10-15.3 (I lcHz spec.) . 10-l3 for 1 s p t 5 30 s.
162 Oscillator hacking 6.4 Quadnoscillators 163

Sp( f), dB rad21Hz 1. The white phase noise is bo = 5.62 x 10-l6 r a d 2 / ~ (-


z 152.5 dB rad2/Hz). Thus

RAKON PHARAO
5 MHz OCXO
assuming that F = 1 dB. Once again, we notice that the power is kept low to achieve
the best stability.
2. The phase flickering is clearly visible in almost two decades, at some 1-100 Hz.
At the corner frequencies f; = 1.5 Hz and f,' = 50 Hz, the difference between the
experimental data and the asymptotic approximation is 3 dB, as expected.
3. The residual noise of the instrument is not shown on the plot. Yet experience
suggests that the value (b- 1),,, = 5.6 x 10-l4 rad2/Hz (- 135.5 dB) is significantly
larger than the noise of a HF-VHF mixer. These elements indicate that the value
(b-l)o,, = 2.82 x 10-l4 rad2/T3z can be trusted.
4. From the plot, the frequency flicker is b-3 = 6.31 x 10-14. We use it for a first
-'
estimate of the Leeson frequency, at which b-3 f = b- f : the result is = 1.5ft
Hz.
5. Let us assume that the sustaining amplifier contributes one-quarter of the total
phase flickering. This is equivalent to guessing that the buffer consists of three
stages (subsection 3.3.2), on the basis that the technology is similar to that of the
Fourier frequency, Hz sustaining amplifier. It follows that

Figure 6.6Phase-noise spectrum of the RAKON PHARAO oven-controlledquartz oscillator.


Courtesy of RAKON (formerly CMAC), 2007. The interpretation, comments, and any mistaltes
are those of the present author. thus f,= 13 Hz.
6. A provisional estimate of the Leeson frequency f{ 2 3 Hz can now be made.
7. Using f : = 3 Hz in fL= vo/(2Q), we get Qs 2: 8.4 x lo5.
RAKON PHARAO 5 MHz quartz oscillator 8. The value Q, 2: 8.4 x lo5 is too low a value for a top-technology 5 MHz quartz
resonator. The quality factor can be 2 x lo6 or more in actual load conditions, that
This oven-controlled quartz oscillator is manufactured in a small series of manually
is, including the dissipative effect of the oscillator circuit.
trimmed samples, intended as the flywheel for the PHARAO atomic fountain. In this
9. Trusting the technology, we use Q, = 2 x lo6. Thus
application, a lower frequency flicker would be required, if available. Figure 6.6 shows
the phase-noise spectrum of one oscillator. This spectrum is obtained as half the raw
spectrum measured with two equal oscillators. In the process of selecting manually the
best samples out of a small series, it is quite reasonable to trust the hypothesis that the
two sections under test are almost equal. This indicates that the oscillator frequency flicker b-3 2 6.3 1 x 10-l4 is due to
The spectrum is again of type 2 (fL < A), typical of low-frequency oscillators with frequency flickering in the resonator, and that the corner frequency f[ 2: 3 Hz of
high-Q resonators. It is fitted by the polynomial c:=-, bi f',with Fig. 6.6 is not the Leeson frequency. Instead, the Leeson effect is hidden by the
resonator's instability and by the phase noise of the output buffer.
10. The flicker noise of the sustaining amplifier combined with the guessed Leeson
bo -152.5 dB 5.62 x 10-l6 rad2/Hz frequency of 1.25 Hz gives a stability limit for the electronics of the oscillator. This
b- 1 -135.5 dB 2.82 x 10-14 rad2/Hz is the solid line f-3 in Fig. 6.6, some 7-8 dB below the phase noise.
b-2 (not visible) 11. The frequency flicker expressed as the Allan variance (Table 1.4), is
b-3 -132 dB 6.31 x 10-14 rad2/Hz

and interpreted as follows.


164 Oscillator hackin 6.4 QuaPtzoscillators 165

thus

crj(r) E 3.5 x and so o,(r) E 5.9 x 10-l4 .


resonator instability
-801; b-, = -1 16.6 dl3 rad2/~z FEMTO-ST prototype
LD cut, 10 MHz, Q = 1.15 x lo6
FEMTO-ST LD-cut quartz oscillator (10 MHz) -90 Leeson effect (hidden)
= -123.2 dB rad2/Hz
-100
Owing to the nonlinearity inherent in a quartz crystal, the resonator's natural frequency
depends on the driving power. This phenomenon, referred to as the amplitude-to- -1 10
frequency effect or sometimes as the isochronism defect, is of order 1.2 x l ~ - ~ / p W
calculated from Av/(voAP) for the popular SC-cut crystals. Attempts to eliminate the
isochronism defect led to the LD cut [42, 351, in which the residual power dependence
is a few parts in I O - ~ ~ I ~ W .
The oscillator analyzed in this subsection (Fig. 6.7) is a laboratory prototype; thus we
lmow two important parameters, the quality factor and the power. This oscillator was I I\',
-160
intended to demonstrate the frequency stability of an LD-cut resonator at low Fourier
\'
1 I i \ 1 4 i I

0.1 1 ; ;o 100 1000 10000 100000


frequencies, in the 11f region. This is unfortunate for our analysis because some pieces
of information were not identified as important and were discarded from the record. In f L = 4.3 Hz 1
I
:f: = 9.3Hz
I
Fourier frequency, Hz
particular, the white-noise level of Fig. 6.7 is inconsistent with the amplifier input power
calculated I : apparent
and with general experience. Keeping this in mind, we proceed with the analysis. Figure 6.7 Phase noise of a FEMTO-ST oscillator prototype based on a LD-cut resonator;
Figure 6.7 shows the phase-noise spectrum, measured as half the total noise obtained Po = 340 pW, the dissipated power = 200 pW. The interpretation, comments, and any mistakes
by comparing two equal oscillators. The plot is fitted by the polynomial c:=-, bi f' are those of the present author.
with
bo -147 dB 2 10-l~ rad2/HZ 5. We use the frequency flicker be, = 2.2 x 10-l2 to estimate the Leeson frequency, at
b- 1 -130 dB 1 10-l3 rad2/~z which b-3 f -3 = b-1 f . A value -'
= 8.9 Hz results. ft
b-2 (not visible) 6. The details of the buffer were unfortunately lost. Hence, our analysis has to rely on
b-3 -1 16.6 dB 2.2 x 10-l2 rad2/Hz the hypothesis that the sustaining amplifier contributes one-quarter of the total phase

-
We interpret the spectrum as follows. flicltering, as in commercial "black boxes." It follows that

1. As we lmow the power at the amplifier input (Po = 340 pW), we can calculate the
noise figure F = bo/(kT). The result is F 167 (22 dB), which is far too high for a
HF-VHF amplifier. Consequently, the measured floor cannot be the noise floor bo of 7. The quality factor of the resonator in actual load conditions is lmown, Q =
the oscillator. Instead, it could be due to insufficient power at the input of the noise 1.15 x 106. The calculated Leeson frequency is
measurement system or to other experimental problems.
2. This oscillator is a laboratory prototype intended to demonstrate the benefit of the
LD cut at low Fourier frequencies. From this standpoint, it is experimentally correct
This value suffers from the uncertainty involved in guessing the phase flicltering of
to ignore a problem showing up in the white-noise region while still trusting the other
the sustaining amplifier as a fraction of the total phase flicltering.
measured noise coefficients.
8. The frequency at which b-31 f = b-1 /f is f[ = 9.3 Hz. This value is significantly
3. The oscillator bo is not accessible, thus the amplifier corner frequency fc cannot be
higher than the calculated Leeson frequency fL = 4.35 Hz. This fact provides evi-
estimated.
dence that the Leeson effect is not visible on the spectrum and that the l lf 3 line is
4. The phase flickering is clearly visible over a span of one decade. At the corner
the frequency flicltering of the resonator.
frequencies f; = 5 Hz and f,' = 50 Hz, the difference between the experimental
9. The frequency flicker expressed as the Allan variance (Table 1.4) is
data and the asymptotic approximation is 3 dB as expected. Experience suggests that
the value (b-l)osc = 10-l3 can be trusted because it is sufficiently large compared
with the noise of the HF-VHF mixer used to measure the oscillator phase noise.
166 Oscillator hacking 6.4 Quarlcz oscillators 167

S,( f), dB r a d 2 / ~ z 2. This oscillator is intended for regular production rather than for special applications.
measurement Agilent 10811 It is therefore reasonable that the phase noise is somewhat higher than that of some
- 7 - system roll off 10 MHz OCXO of the special units already discussed. This fact makes its measurement easier, which
further encourages us to trust the spectrum of Fig. 6.8.
o experimental data 3. At the corner frequencies, 25 Hz and 1.3l a z , the difference between the experimental
sust. ampli. + buf. data and the asymptotic approximation is 3 dB, as expected.
b-I = -131 dB 4. In the 50-300 Hz region, the phase noise is of the l / f type. This includes sustaining
amplifier and buffer. Taking away 6 dB (three buffer stages), the estimated flickering
of the sustaining amplifier is

thus f, = 320 Hz.


5. The oscillator l/f noise crosses the sustaining-amplifier noise (b-i)ampli f at -'
f[ -.50 Hz. Whether f{ is the Leeson frequency is still to be decided.
6. Using f; = 50 Hz in fL = vo/(2Q), we get Qs 2. 1 x 10'.
fL=:7 Hz1 7. The value Qs E 1 x lo5 is far too low for it to be the resonator quality factor. To
(guess !)
f: = 50 HZ fc =' 320 Hz frequency, Hz be conservative, we believe that the quality factor Qt cannot be lower than 7 x 105.
Thus
Figure 6.8 Phase noise of the Agilent 10811 10 MHz oscillator [15]. Used with the permission of
Agilent Technologies, 2007. The interpretation, comments, and any mistakes are those of the
present author.

hence Accordingly, the oscillator frequency flicker bw32 5 x 10-I is due to the frequency
oi(r) 2 3 x and so oy(r) -. 1.74 x lo-" . flickering in the resonator, while the Leeson effect is some 17 dB lower.
8. The frequency flicker expressed as the Allan variance (Table 1.4) is
Agilent 10811 quartz (10 MHz)
b-3 5 x lo-"
o;(r) =21112h-~ =2ln2, = 1.39 x
This oven-controlled quartz oscillator has been the high-stability workhorse for a number vo (10 x 106)2 '
of Agilent - and formerly Hewlett Pacltard - electronic instruments. It is no longer sup-
ported by Agilent but the manual is available on a website maintained by the community thus
of users [5 11.
Figure 6.8 shows the phase-noise spectrum. The spectrum is fitted by the polynomial oi(r) E 6.9 x loe2' and so oJ,(r) -. 8.3 x 10-l3 .
c:=-, bi f', with
The 1/f noise of the sustaining amplifier deserves a final comment. This noise is inferred
bo -162 dB 6.3 x 10-17 rad2/Hz by subtracting 6 dB from the I/ f region of the output spectrum, in the same way as
b- 1 -131 dB 7.9 x 10-13 rad2/Hz for other oscillators analyzed in this chapter. It is surprising, however, to notice that the
b-2 (not visible) numerical value of this 11f noise is similar to that of the other oscillators, most of which
b-3 -103 dB 5.0 x lo-" rad2/Hz are intended for far more demanding - and far more expensive - applications.
We interpret the spectrum as follows.
1. Under the usual assumption that F = 1 dB, the white phase noise yields Agilent noise-degeneration oscillator (10 MHz)
This oscillator, tested at the Agilent Laboratories, [60], includes a noise-degeneration
system based on the ideas of subsection 2.5.3, which reduces the flicker noise of
The moderate operating power helps in achieving high long-term stability. the amplifier. Figure 6.9 shows the phase-noise spectrum. The plot is fitted by the
Oscillator hacking 6.4 Quarliz oscillators 169

3. The visible corner frequency, at which the f -3 noise crosses the f -' noise, is
-80
Agilent prototype
b-3 = -102 dB rad2/Hz 10 MHz OCXO
1 4. Subsection 3.3.2 indicates that the effect of the output buffer cannot be neglected if
sust. arnpli. + buf. phase flicltering shows up in the phase-noise spectrum. Elsewhere we have assumed
b-l = -126 dB rad2/Hz
that there are three buffer stages between the sustaining amplifier and the output,
simple sust. amplifies the I/ f noise of each stage being equal to that of the sustaining amplifier and the
b - -132
-zlZ- dB r a d 2 / ~ z
-----r -- fliclter noise of the sustaining amplifier being (b-l)an,pa = 1 (b-l)tot. Thus
I bo=-158dBrad2/Hz
I I

(for a simple amplifier, not for this oscillator) .

This is indicated as a broken line in Fig. 6.9.


5. Owing to feedback, the noise-degeneration amplifier exhibits low flicker, limited by
the internal phase measurement system. Let us assume that the 1/f noise improves
I
I'
I Fourier frequency, Hz on a non-corrected amplifier by 20 dB, as a conservative value. Hence, an upper
fc=4HzI f;=320Hz bound of the amplifier noise is
I
.- - - - - _ _ _ _ - _'corrected
-_ sust. amplifier '
fL z 7 Hz (guess!)
b-l = -152 d~ rad2/Hz
L

(noise corrected amplifier) .


Figure 6.9 Phase noise of the balanced-bridge quartz oscillator tested at Agilent Laboratories
(601. Used with the permission of the IEEE, 2007. The interpretation, comments, and any In Fig. 6.9, this is the solid line 20 dB lower than the noise of a simple amplifier and
mistakes are those of the present author.
26 dB lower than the oscillator output 1/f noise.
6. The corner frequency of the sustaining amplifier is
polynomial c:=-, bi f',with

bo -158 dB 1.6 x 10-l6 rad2/Hz


b-1 -126 dB 2.5 x 10-l3 rad2/Hz 7. The visible f -3 noise crosses the estimated f -' noise of the sustaining amplifier at
b-2 (not visible)
b-3 -102 dB 6.3 x 10-l1 rad2/Hz
z 320 H z .
We interpret the spectrum as follows.

1. The noise-degeneration amplifier exhibits low fliclter, at the expense of a higher 8. Taking f[ = 320 Hz as the Leeson frequency and using fL = vo/(2Q) we get
noise figure than that of a simple amplifier. This higher noise figure is due to the Qs E 1.6 x lo4.
need to split the input signal into two branches, for power amplification and for 9. The value Qs z 1.6 x lo4 is far too low a value for a state-of-the-art oscillator. The
noise correction of the power amplifier (subsection 2.5.3). We assume that F = 5 literature suggests that the product voQ is between 1013 and 2 x loL3(unloaded),
dB; this is inferred by accounting for the 3 dB loss inherent in power splitting, plus reduced by a factor 0.5-0.8 when loaded.
the 1 dB dissipative loss and 1 dB noise figure of the internal amplifiers. 10. The actual Q value is higher than 1.6 x lo4. This means that the Leeson effect is
2. The white phase noise is bo = 1.6 x 1 0 1 6 rad2/Hz (- 158 dB rad2/Hz). Thus hidden, and that we are not able to calculate Q. Hence we can only guess the value
of Q, relying on experience and on the literature.
11. If we admit Q = lo6, reduced to Q = 7 x lo5 (loaded) by the dissipative effect of
the amplifier input, we get
This power level is consistent with one's general experience of 5-10 MHz oscillators,
in which the power is ltept low for best stability.
170 Oscillator hacking 6.4 Quarltz oscillators 171

This would indicate that the oscillator frequency flicker b-3 2 6.3 x 10-I is due to -100
-30 dB/dec VVenzel501-04623
frequency flickering in the resonator, not the Leeson effect. Accordingly, the corner b-, = -67 dB rad2/kIz 100 MHz OCXO
visible in Fig. 6.9 at 16 Hz will not occur at the Leeson frequency.
12. Using + specifications

we find that the contribution of the Leeson effect to the oscillator l /f noise is

This value is about 33 dB lower than the actual b-3 coefficient of the oscillator
noise.
13. With reference to Fig. 3.12, we note that the spectrum is of type 1 B, for which
fc c fL. It is worth mentioning that we have chosen an upper bound for the amplifier
l / f noise. A value lower than - 152 dB r a d 2 / ~ results
z in a value off, lower than
4 Hz, which reinforces the conclusion that f, ifL.
14. The frequency flicker expressed as the Allan variance (Table 1.4), is
101 102 ;I 103 fL= 3.5 kHz
104 105
guess Q = 04 =>fL.= 625 Hz Fourier frequency, Hz

Figure 6.1 0 Phase noise of the 100 MHz oscillator Wenzel50 1-04623. The numerical values are
thus taken from the manufacturer's website. The plot, interpretation, comments, and any mistaltes are
those of the present author.
o ; ( r ) ~ 8 . 8 ~ 1 0 - ~ ~andso 0~,(~)~9.3~10-'~,
Wenzel501-04623 (100 MHz SC-cut quartz)
There is some disagreement between this value and the value o,, = 2 x 10-l2 at
r = 1 s given by [60, Table I]. This disagreement is present in the article cited and The Wenzel 501-04623 oscillator is chosen as an example because of its outstandingly
is independent of our interpretation. low phase-noise floor and because of its high carrier frequency (100 MHz). These two
15. The Allan variance calculated only on the basis of the Leeson effect, thus with a features are necessary for low noise after frequency multiplication. A table of phase-
noise-free resonator but accounting for an upper bound of the amplifier 1/f noise, noise values is given by the manufacturer, rather than the complete spectrum. These
is values, used in plotting the spectrum (Fig. 6. lo), are:
bo -173 dB 5 x lo-'' rad2/Hz
b-1 (not visible)
b-2 (not visible)
thus b-3 -67 dB 2 x lo-7 rad2/Hz
o;,(r) 2 2.1 x 10-l4 (noise-free resonator) . An interpretation is as follows.
Of course, this high stability is out of reach for a quartz oscillator. 1. The white phase noise is bo = 5 x 10-l7 r a d 2 / ~ (-
z 173 dB rad2/Hz). Thus
Flc T
When this project was presented at the 1999 EFTFIFCS joint meeting in Besan~onand Po = -~ 1 m W (OdBrn),
published in [60], a number of researchers and engineers then expected a new low-noise bo
oscillator from Agilent. This did not happen. The reasons are of course confidential. assuming that F = 1 dB. This relatively large power is necessary for low short-term
Nonetheless, we can infer that the new and smart sustaining amplifier was an unnecessary noise, although it may be detrimental to the medium-term and long-term stability.
improvement on a well-designed traditional amplifier, since the oscillator stability is in -'
2. The terms b-2 f -2 and b-t f must be negligible because there is no room for them
any case limited by the fluctuations of the resonator's natural frequency. in the region between the two specified points, 1 Id-Iz and 10 kHz. Confirmation
172 Oscillator hacking 6.5 "The origin of instabiliw in quarZz oscillators

comes from the measurement of some samples. Consequently, the point at which the uncertainty in identifying or guessing f:. Additionally R is related to the quality factor
slope changes by 2 (either f O to f -2 or f-I to f-3) is not visible. This indicates that by
the Leeson effect is hidden by the frequency flicker of the resonator.
3. Of course, the corner point where the frequency flicker crosses the white noise R % -Q 25
Qt
---
(b-3 f -3 = bo), which occurs at f = 3 lcHz, is not the Leeson frequency. Qs Qs'
4. A further consequence of the absence of a visible 1/f region in the spectrum is that we
where Qs is calculated using f[ as if it were the Leeson frequency and where the true
have no data from which to infer the 1/f noise of the sustaining amplifier. The flicker
Q is replaced by Qt, that is, the most plausible value based on the available technology.
noise of a state-of-the-art HF-VHF amplifier is between - 135 and - 140 dB r a d 2 / ~ z .
This can be seen by introducing the power law (1.70) into the Leeson formula, written in
Thus, we will take (b-l)aqii 1.6 x 10-l4 r a d 2 / ~ (-
r 138 dB r a d 2 / ~ z as
) a plau-
terms of vo and Q, (3.19), or in terms of fL, (3.2 1). The above-mentioned approximation
sible value.
derives from f,: which enters in Qs, and then from Qt.
5. As the Leeson effect is hidden, we are not able to calculate Q. Relying upon the
Finally, R can be expressed as the stability ratio
literature, we guess that a high-stability 100 MHz quartz resonator will have a quality
factor of 1.2 x 105, reduced to Q = 8 x 1o4 in actual load conditions. Accordingly, (aJ,)osc
R = ------ (flicker floor) ,
the Leeson frequency is
(OJ,)L

that is, the Allan-deviation floor of the complete oscillator calculated from the observed
1/f phase noise, divided by the floor calculated on the basis of the Leeson effect only,
6. The sustaining-amplifier l / f noise, combined with the Leeson frequency, gives the
assuming that the resonator is free from noise. Equation (6.10) is derived from (6.7)
stability limit of the electronics of the oscillator. This is the broken line 6.2 x loF9/f
using Table 1.4 or Fig. 1.8. The parameter R states the poorness of the actual oscillator,
in Fig. 6.10 that is 15.1 dB below the phase noise.
as compared with the same oscillator if it were governed only by the Leeson effect
7. The frequency flicker expressed as the Allan variance (Table 1.4), is
and had no resonator fluctuations. This statement can be rephrased by saying that R
is the goodness of the sustaining amplifier compared with the oscillator fluctuations.
Thus:
thus
R = 1 (0 dB) indicates that the oscillator f -3 phase noise comes from the Leeson
o j ( r ) EZ 2.8 x and so o;,(r) E 5.3 x 10-l2 . effect and that the resonator fluctuations are negligible;
R = f i (3 dB) indicates that the resonator l / f fluctuations and the Leeson effect
contribute equally to the oscillator noise;
R >> f i (3 dB) indicates that resonator instability is the main cause of the f -3
6.5 The origin of instability in quartz oscillators phase noise.

Let us introduce the noise ratio R, defined as Table 6.1 gives a comparison of the results. Some 5-10 MHz oscillators are similar,
to the extent that all the available technology is employed in order to get the highest
stability. In all cases we have analyzed, we find R values of the order of 10 dB, with a
minimum of 6.6 dB. This means that the Leeson effect is hidden below the frequency
This is the total frequency fliclcer at the amplifier output divided by the frequency fliclcer fluctuation of the resonator. The Agilent 10811 is closer to routine in production, and
due to the sustaining amplifier via the Leeson effect. Although not mentioned explicitly, probably closer to a cost-performance trade-off, than the others; thus understanding this
R can be easily identified in all the phase-noise spectra of this chapter. As a consequence oscillator is more difficult. Nonetheless, in this case the value of Qs is so low that there
of (6.7), the parameter R is also given by is no doubt that it cannot be the resonator's quality factor. The Wenzel 50 1-04623 (100
MHz) is clearly designed for minimal phase noise in the white region. This requires a
high driving power, so it is not surprising that the resonator stability is impaired.
The examples shown above indicate that an oscillator's frequency Aicltering is chiefly
where f[ is the frequency at which the continuation of the oscillator l / f noise crosses due to the fluctuations in the resonator's natural frequency. This fact has a number of
the 1/f noise of the sustaining amplifier (Fig. 6.2). The approximation is due to the implications, some of which are unexpected.
6.6 Microwave oscillators 175

1. An oscillator7sfrequency flicltering can only be improved by improving the resonator.


2. As a consequence, there is no way of improving the frequency flicltering through
innovative design of the sustaining amplifier. To this extent, the electronics is at a
dead end.
3. The analysis of the 10811 oscillator confirms the above conclusion. As a matter
of fact, the noise of the sustaining amplifier is similar to that of higher-stability
oscillators. We do not laow anything about the cost-performance trade-off in the
design of this oscillator. Nonetheless, we notice that the I/ f noise of the sustaining
amplifier is not increased by a lower budget.
4. Surprisingly, in ultra-stable oscillators the 1/f phase noise of the output buffer shows
up, and perhaps also the 1/f noise of the sustaining amplifier, which can be annoying
in the 0.1-1 s region of the Allan deviation. This fact leaves room for innovative
electronic design.
5. After observing that s'mple and relatively cheap radio-electronics is sufficient for
the stability of a high-end oscillator to be limited by the resonator, it may be
pointed out that the stability of lower-rank oscillators is clearly dominated by the
resonator.
6. Finally, we recall that this analysis is based only on the phase noise affecting the
amplifier's forward gain. The noise originating in the impedance fluctuation of the
amplifier transferred to the resonator's natural frequency (subsection 3.3.4) has not
been talten into account. This noise phenomenon, whose existence is even not men-
tioned in the literature, deserves theoretical development and experiments.

6.6 Microwave os~illatovs

Microwave oscillators differ radically from quartz oscillators in that the 1/f frequency
noise is governed by the amplifier's I / f phase noise through the Leeson effect. This is
related to three differences between microwave oscillators and quartz oscillators.

1. The volume of energy confinement is significantly larger in microwave resonators


than in quartz resonators.
2. Excluding cryogenic units, the quality factor of a microwave resonator is usually
lower by a factor 10-lo3 than that of a quartz resonator.
3. Microwave amplifiers show larger 1/f noise than W amplifiers.

Additionally, circulators and isolators are available only at microwave frequencies, while
the reverse isolation of a microwave amplifier is usually rather poor.
This section gives a number of worked-out examples of microwave oscillators.

Miteq DRO mod. D-21OB


Figure 6.1 1 shows the phase-noise spectrum of the dielectric resonator oscillator
(DRO) Miteq D-210B, talten from the device data sheet. The plot is fitted by the
Oscillator hacking 6.6 Microwave oscillators 177

3. The spectrum changes slope from f O to f -2 at the Leeson frequency fL 2i 4.3 MHz.
+3
-7 S,( f ), dB rad2/Hz 10 GHz DRO Miteq D210B
At this frequency, the asymptotic approximation is some 4 dB lower than the measured
-17 spectrum, instead of the expected 3 dB. This discrepancy is tolerable.
4. From fL 2i 4.3 MHz, it follows that Q = v0/(2fL) 2i 1160, which is quite plausible
for a dielectric resonator.
5. The corner point at which the slope changes from -2 to -3 is 70 1cHz. This is the
corner frequency f, of the amplifier, at which it holds that (b-l)ampli -'
f = (bo)ampli.
Hence (b-l)ampli= 1.8 x 10-lo rad2/Hz (-98 dB r a d 2 / ~ z ) .
6. The flicker-frequency coefficient b-3 = 5 x 1o3 rad2/Hz (+37 dB rad2/Hz).
7. The white and fliclter frequency noise, transformed into the Allan variance (Table 1.4),
(b-l)ampli= -98 dB sad2/~z 4 dB difference is

j
I

103 104 105 106 1 107


: Fourier frequency, Hz I

fc = 70 kHz f, = 4.3 MHz

Figure 6.11 Phase noise of the 10 GHz DRO Miteq D21OB. Used with the permission of Miteq
Inc., 2007. The interpretation, comments, and any mistakes are those of the present author.
thus

polynomial c!=-, bi f',with

bo -146 dB 2.5 x lo-'' rad2/Hz


b- 1 (not visible)
b-2 -11 dB 7.9 x 1oY2 rad2/~z
b-3 +37 dB 5.0 x lo3 rad2/HZ

This indicates that the spectrum is of the type 1A of Fig. 3.12. Finally, one should note that the oscillator fliclter shows up in the 1-100 kHz region.
One might be tempted to fit the spectrum with a smaller bo (say, - 147 dB rad2/Hz) and Common sense suggests that temperature and other environmental fluctuations have no
-'
to add a term b- f tangent to the curve at f FZ 2 MHz. In this case the spectrum would effect on this time scale and that the flickering of the dielectric constant in the resonator
be of type lB, which contains the signature of the output buffer. We will discard this will not exceed the amplifier noise. Consequently, in this region the oscillator flicker is
interpretation, however, because then the noise of the output buffer would be b- 1 % 1o - ~ due to the amplifier, through the Leeson effect, rather than to the resonator.
rad2/Hz (-80 dB rad2/Hz), which is too high for a microwave amplifier (Table 2.1).
The spectrum gives the following indications.
Poseidon DRO-10.4-FR (10.4 GHz)
1. The coefficient bo derives from the amplifier noise FliT referred to the carrier power The Poseidon DRO-10.4-FR is another example of an oscillator based on a dielectric
Po at the input of the sustaining amplifier, that is, bo = FkT/Po. Assuming that the resonator. Figure 6.12 shows the phase-noise spectrum, from a preliminary data sheet.
noise figure is F = 1 dB, and thus FliT = 5.1 x 1o - ~rad2/Hz (- 173 dB rad2/Hz), The unusually low noise raises the question whether this DRO is a technology demon-
it follows that Po = FliT/bo = 2 pW (-27 dBm). strator or a regular product. The power spectral density S,( f ) is fitted by the polynomial
2. However arbitrary the assumption F = 1 dB may seem, it is representative of actual ~ f = bi-f ~, with
microwave amplifiers. Depending on bandwidth and technology, the noise figure of
a "good" amplifier is between 0.5 dB and 2 dB. In this range, we find Pobetween
1.8 pW and 2.5 wW. The symbol x means us~~r?ptoticully
equal.
178 Oscillator hacking 6.6 Microwave oscillators 179

6. The corner frequency of the amplifier (i.e. the frequency at which the oscillator
spectrum changes from f -2 to f -3) is fc = 9.3 1cHz. Accordingly, the phase-noise
-57 Two Poseidon DRO-10.4 oscillators spectrum of the amplifier, on the left-hand side of f = fc, is (b-l)ampli= bOfc=
-67 2.9 x 10-l3 rad2/Hz (- 125 dB rad2/Hz).
7. The amplifier flicltering, 5 dB lower than the best value of Table 2.1, is surprisingly
low for a microwave amplifier within a commercial product. Such a low noise could

.7
slo e close to -25 dB/dec
be obtained from SiGe technology, with a single-stage amplifier employing a large-
volume transistor, or from a feedback or feedforward noise-degeneration scheme.
A noise-degeneration scheme seems incompatible with the size of the packaged
oscillator, yet nothing definite can be inferred on the basis of the available information.
8. The fliclter frequency coefficient is b-3 = 2.5 r a d 2 / ~ (+4
z dB rad2/Hz).
9. The white and fliclter frequency noise, transformed into the Allan variance (Table 1.4),
is

102 103 :104


I
105 lo6 :
I
107
Fourier frequency, Hz
f,= 9.3 kHz f, = 3.2 MHZ
Figure 6.12 Poseidon DRO 10.4-FR.Used with the permission of Poseidon Scientific
Instruments, 2007. The interpretation, comments and any mistakes are those of the present
author.
thus
bo -165 dB 3.2 x 10-17 rad2/Hz
b-1 (not visible)
b-2 -35 dB 3 . 2 ~ 1 0 ~ rad2/Hz
~
b-3 +4 dB 2.5 rad2/HZ
and so
Once again the spectrum is of the type 1A (Fig. 3.12) typical of microwave oscillators,
Yet the discrepancy with respect to the theoretical model is larger than in the case of the
Miteq oscillator. The spectrum gives the following indications.

1. According to the manufacturer's data sheet, the spectrum results from measurements Figure 6.12 also reports the phase-noise spectrum of the DRO- 10.4-XPL oscillator,
on two DRO- 10.4 oscillators. It is common practice to compare two oscillators in this which is a different version of the same base design intended for phase-loclted loops.
way and to refer the spectrum to a single oscillator after taking away 3 dB from the raw Below a cutoff frequency of about 70 IcHz, this oscillator is loclted to an external
data. This is legitimate under the quite realistic assumption that the two oscillators reference. In this condition, the spectrum is chiefly determined by the external reference
are equal and independent and thus that each contributes half the measured spectrum. and gives little information on the oscillator. Nonetheless, it is to be noted that the
2. The white phase noise is bo = 3.2 x 10-l7 rad2/Hz (- 165 dB rad2/Hz). Thus spectrum is proportional to f -5'2 (i.e., -25 dB/decade) below 10 kHz, the loop cutoff
Po = FkT/bo 2i 160 pW (-8 dBm), assuming that I: = 1 dB. frequency. This might be the result of an experimental problem in the measurement or
3. The Leeson frequency is fL = 3.2 MHz. Accordingly, the quality factor the signature of a fractional-order control system, like that proposed in [24].
Q = v0/(2 fL) 2i 1625, which is reasonable for a dielectric resonator.
-'
4. In a type-1A spectrum it holds that b-2 f = bo at f = fL. Thus, the white frequency
Poseidon Shoebox (10 GHz sapphire resonator)
noise is b-2 2i 3.2 x lop4 rad2/Hz (-35 dB rad2/Hz).
5. There is some discrepancy between the Leeson model and the true spectrum. In the The Poseidon Shoebox maltes use of a sapphire whispering-gallery (WG) resonator and
region from 2 kHz to 200 kHz the spectrum seems to be close to a line of slope a bridge noise-degeneration amplifier. The result is a low-noise oscillator intended for
f -5/2 rather than f -3 or f -2. At the present time this discrepancy (up to 4 dB at high short-term-stability applications. Valuable information about this type of oscillator
f 10 1d-I~)is unexplained. or resonator is found in [67] and 1591.
Oscillator hacking 6.6 Microwave oscillators 181

resonator size. Moreover, the oscillator size and weight (3 dm3 and less than 7 kg) are
. . . . . . . . incompatible with the presence of a cryo-generator. Finally, such an optical cooling
-87 .. .. ,.. .. .. . .. . .., ... , ... .. .. .. .. .. .. .. ..
,

, .
,

, , , , , a , ,
,
,
, . , , , , ,
. , , , , , Poseidon Shoebox : I technology [30, 281 is still not available and was definitely out of question when this
oscillator was designed. These issues indicate that the resonator is intended to be
used at room temperature.
3. The phase flicker of the sustaining amplifier is b-l f with b-1 w 10-I2 r a d 2 / ~ z . -'
What is seen on the plot is - 160 dB r a d 2 / ~ atz f = 10 liHz with slope f- l cor-
responding to -120 dB r a d 2 / ~ zat 1 Hz. This value is too high for a sophisticated
amplifier that maltes use of the bridge noise-correction technique.
4. At f = fL, where b-3 f -3 = b-l f the true spectrum differs from the asymptotic-',
approximation by 6 dB instead of 3 dB. This discrepancy is an additional reason to
reject the interpretation of Fig. 6.13.

The above difficulties lead 11s to understand that the spectrum is actually of type 2A
(Fig. 3.12), in which the fliclter noise of the output buffer shows up in the region around
fL. There follows a completely new interpretation, shown in Fig. 6.14.
Fourier frequency, Hz

Figure 6.13 First attempt to interpret the phase-noise spectrum of the Poseidon Shoebox. Used
1. As usual, we start from the white noise floor. On the figure we observe that
with the permission of Poseidon Scientific Instruments, 2007. The interpretation, comments, bo = 1.3 x 10-l7 r a d 2 / ~ z(-169 dB r a d 2 / ~ z )This
. is ascribed to the sustaining
and mistalces are those of the present author. amplifier.
2. The sustaining amplifier maltes use of a bridge noise-degeneration circuit to reduce
the fliclter noise. In this circuit there are two amplifiers: the first amplifies the input
Figure 6.13 shows the phase-noise spectrum with a tentative interpretation. The spec-
signal and the second is used to correct the noise of the first, at the output of a
trum seems to be of type 2 (Fig. 3.1 I), with fL < f,. Qualitatively, this is consistent with
the fact that the WG resonator features a high Q value. Yet this interpretation suffers
from three problems.
-87 .. . ,. . , . , . , ., , . , ,. ., .. ., ., ,, ., ,, ., ., . . , , . , ,
1. From Q = vO/(2fL) we get Q e 1.9 x lo6 at vo = 10 GHz. This value is incompat-
,

... .. .,. .. .. . .. , .., ,.. , ... .,. .. ... .. , .. , .. . .., , .. . , .. Poseidon Shoebox
ible with the dielectric loss of the sapphire. For comparison, the typical quality factor -97 - . . . . . . . . . ., ..........
:
. .. . ,
% .
.................................
. . . . . :, . :, . :. .,. ., > . . . . . . . . ..,. . . . . . .., . . . .,,. . . ... . .,,. .,,. . ... . .,
...........
.. ,, ,, ., ,, ,, .. .. .. .. .. .. .. .. ..
, , ,
. . , . , . . , , . .

of a 10 GHz WG resonator of 5 cm diameter and 2.5-3 cm thiclcness is given in the


following table:

Q Temperature

2 x lo5 295 I<, room temperature


3 x lo7 77 IS, liquid N
4 x 1o9 4.2 IS, liquid He

. . ,. . . . , . , .. , . . . . . . . . . . . .
The resonator size cannot be reduced significantly without impairing the quality -177 ! .
,
.. .
, .
. : : : I I : :' . . . . . . . . .
,
I : II : : : : : : :
factor. The dielectric loss of the sapphire is a reproducible function of temperature. 100 ;loo0 10000 I
I 100000
Thus, the resonator loss depends on the space distribution of the electric field, i.e., I Fourier frequency, Hz '
f'= 850 Hz fL= 25 lcHz
on the mode, in a narrow range.
2. Cooling to about 50 IS below room temperature is possible with Peltier cells. Yet a Figure 6.14 Phase noise of the Poseidon Shoebox. The same spectrum as that of Fig. 6.13 but
quality factor of 1.9 x lo6 cannot be obtained with this moderate cooling. A power with a revised interpretation. Used with the permission of Poseidon Scientific Instruments,
consumption of 35 W (steady state, at 25 "C) is too low for this technology and 2007. The interpretation, conunents, and any mistalces are those of the present author.
182 Oscillator hacking 6.6 Microwave oscillators 183

carrier-suppression circuit. It does this by means of a feedback circuit. We estimate Prototype 9 GNz whispering-gallery oscillator (liquid-N,77 K)
a noise figure F = 5 dB, which results from the intrinsic loss of the power splitter at -77y +m7. . 1*17CF"-*
I 1 4 1 ) 1 1
I il I * q
the input of the noise-corrected amplifier (3 dB), from the resistive loss of the power
splitter and of the lines (1 dB), and from the noise figure of the second amplifier
(1 dB).
3. From bo = FkT/Po we get Po = 1 mW (0 dBm).
4. The phase flickering of the output buffer shows up on the right-hand part of the
spectrum, at lo5 Hz and beyond. The noise coefficient is (b- 2 10-l2 rad2/Hz
(- 120 dB rad2/Hz). The output buffer is a good microwave amplifier.
5. After removing the buffer phase noise, the white frequency noise b-2 f -2 of the
oscillator may be clearly identified. The decade centered at 4 ltHz is used to find
b-2 = 7.9 x rad2/Hz (-8 1 dB rad2/Hz) on the plot.
6. From b-2 f -2 = bo, found on the plot, it follows that fL 2 25 1cHz. Hence,
Q = vO(2fL) 2 2 x lo5. This is the typically good value obtained for a room-
temperature WG sapphire resonator.
7. It is seen on the plot that b-3 2 7.9 x lo-' r a d 2 / ~ (-51
z dB rad2/Hz).
8. In a spectrum of type 2A, the corner frequency f, of the sustaining amplifier is
the frequency at which the oscillator noise changes slope from f -3 to f -2. Thus,
fc 2 850 Hz. The flicker noise of the sustaining amplifier is (b-l)ampli f-' with Figure 6.15 Liquid-N whispering-gallery 9 GHz oscillator tested at the University of Western
(b-l)ampli= bof,. Thus (b-l)aqli E 10-l4 r a d 2 / ~ (- z 140 dB rad2/Hz). Australia [108]. Used with the permission of the IEEE, 2007. The interpretation, comments, and
9. The white and flicker frequency noise, transformed into the Allan variance any mistakes are those of the present author.
(Table 1.4), is
temperature to 3 x lo7 (unloaded); the thermal coefficient of the resonant frequency
decreases from -7 x 10-~/1<to - 1 x 10-5/~<;and the thermal conductivity increases
from 24 W m-'1<-l to 960 W m-l~c-l, which ensures temperature uniformity and in
turn reduces the mechanical stress. The two reference boolts on these types of oscillator
and resonator are [67] and [59].
Figure 6.15 shows the phase noise of a prototype 9 GHz oscillator built around a liquid-
N whispering-gallery resonator, tested at the University of Western Australia [lOS].
thus This oscillator includes a noise-degeneration amplifier based on the ideas discussed in
subsection 2.5.3. The phase-noise measurement of this type of oscillator is difficult,
and for this reason [lo81 gives the result as a set of fragments of information instead
of as a single phase-noise spectrum. Thus, we will analyze only the oscillator with the
and so noise-degeneration loop disabled. The phase-noise spectrum is curve a of Fig. 6.15. This
plot is fitted by the polynomial c:=-, bi f ' , with

bo -136 dB 2.5 x 10-14 rad2/Hz


10. The bump at f = 40 kHz is ascribed to the feedback control of the noise degener- b- 1 -101 dB 8x10-l3 rad2/Hz
ation circuit. b-2 (not visible)
b-3 -50 dB 1 x loy5 rad2/HZ
UPlVa liquid-l\l whispering-gallery 9 GHz oscillator We interpret the noise spectrum as follows.
When cooled to liquid-N temperature (77 I<), a sapphire resonator improves signifi- -'
1. The f and f -3 noise types show up clearly, while there is no f -'noise. Hence,
cantly in all the relevant parameters. The quality factor increases from 2 x lo5 at room the spectrum is of type 2 (fL < f,).
184 Oscillator hacking 6.7 Optoelectronic oscillators 185

2. The white noise is only partially visible. This occurs because the oscillator is 6.7 Optoelectronic oscillators
measured by comparison with a discriminator, which is a resonator of the same
type as that in the loop (Fig. 5 of [108]). Away from the discriminator bandwidth
the output signal loses power; thus the phase detector loses gain and the voltage NIST 10 GHz opto-electronic oscillator (OEO)
noise at the detector output decreases. This leads to the wrong conclusion that the The unique feature of the optoelectronic oscillator (OEO), developed at the Jet Propulsion
noise spectral density still decreases beyond 20-30 16-I~.Trusting the theory, we Laboratory (JPL) [I101 in the mid 1990s, is that the frequency-selective element is an
find bo using two criteria: optical delay line instead of a traditional resonator. The basic structure consists of a loop
the 3 dB difference between the straight-line approximation and the true spectrum; in which the microwave sinusoid modulates a laser beam. After traveling through a long
the value in the narrow, nearly flat, region at f 10 kHz. optical fiber, the microwave signal is detected by a high-speed photodetector and fed
3. The noise-degeneration circuit, although disabled, is still present in the circuit and back into the intensity modulator after amplification.
leads to an intrinsic and dissipative loss. For this reason, we assume that the noise We analyze the noise of a prototype tested at the US National Institute of Standards
figure F is 5 dB. and Technology (NIST) and published in the article [78]. The reason for choosing this
4. The white phase noise bo = 2.5 x 10-l4 rad2/Hz. Thus Po = FkT/bo = 5 x oscillator is that the published article, as compared with other references, discloses more
W (-33 dB m), with F = 5 dB. technical information specific to the prototype. The phase-noise spectrum, shown in
-'
5. The corner point at which b-1 f = bo occurs at fc = 3.1 16-I~.The scheme of the
oscillator (Fig. 3 of [1OX]) indicates that there is no output buffer. Hence the b- f-'
Fig. 6.16, is fitted by the polynomial c:=-, bi f',with

region of the spectrum is due to the flickering of the sustaining amplifier. The value Sqmin -137 dB 2 x 10-14 rad2/~z
- 101 dB is plausible for a microwave amplifier. b- 1 (not visible)
6. According to the scheme published, the oscillator is followed by two isolators and no b-2 -42 dB 6.3 x 1o - ~ rad2/Hz
buffer. With this configuration, isolation would probably be insufficient for practical b-3 -10 dB 0.1 rad2/Hz
applications. Yet this can be tolerated in a laboratory environment, with the oscillator b-4 +2 dB 1.6 rad2/Hz
connected to a stable load. We interpret the spectrum as follows.
7. The frequency at which b-3 f -3 = bP1f -', is fL = 360 Hz. In the type-2 spectrum
we have no means of saying whether this corner point is the true Leeson frequency 1. In the case of an OEO, the "sustaining amplifier" also includes the optical modulator
rather than an effect of the fluctuating frequency of the resonator. Better a priori and the photodetector. For short, the subscript 'ampli' will be used for notational
lcnowledge of the loaded quality factor is necessary, which we do not have. However, consistency, instead of making explicit reference to the above chain of devices. The
in this case we can find the answer somewhere else. In fact the f -3 noise improves structure of this chain will be introduced when needed.
significantly after introduction of the noise-degeneration circuit, since the resonator
l / f 3 noise is lower than that of plot a in Fig. 6.15. Consequently, fL = 360 Hz must
be the true Leeson frequency.
8. With fL = 360 Hz, we find Q = vo/fL = 1.25 x 10'. This is a reasonable value for
the loaded quality factor of this type of resonator.
9. There is remarltable agreement between the two estimates of fL, that based on the
crossing point of the straight lines and that based on the 3 dB difference between
the straight lines and the measured spectrum.
10. The frequency flicker transformed into the Allan variance (Table 1.4) is

thus

Fourier frequency, Hz
and so Figure 6.16 Phase noise of the NIST OEO prototype, employing 1.2 lun optical fiber as a delay
line. The spectrum shown is Fig. 7 of [78], used with the permission of the IEEE, 2007. The
interpretation, comments, and any mistakes are those of the present author.
186 Oscillator hacking

2. From [78], the length of the optical fiber is L = 1.2 lun. With a typical refraction
index of 1.468, the expected delay is t d = 5.88 ps. Thus the base frequency must
be l / t d = 170 kHz.
3. From the phase-noise transfer function H(s), (5.69) (also shown in Fig. 5.14),
we expect a minimum at f E 1/(2td) 85 kHz. At this minimum, it holds that
s , ~ = ) as$(f).
4. The sequence of pealts and minima is not visible in Fig. 6.16 because the frequency
axis is truncated at 100 kHz. Only the minimum at 85 l H z is visible.
5. The value tabulated as SPmin = 2 x 10-l4 rad2/Hz (- 137 dB rad2/Hz) is not the
floor bo of the oscillator. It is the first minimum predicted by (5.69).
6. It is seen in Fig. 6.16 that the corner frequency at which the l / f noise turns into
11f noise occurs at 1.6 kHz. This corresponds to the corner frequency fc of the
electronics, amplifier, and photodetector, where the noise turns from f O to l / f .
7. The minimum at 85 kHz is in the white-noise region of the electronics, where the
corner frequency fc is far enough away for the l / f noise to be negligible. Thus,
using S,(f) = aS$(f) we get
Figure 6.17 Phase-noise model of the the NIST OEO prototype (Fig. 6.16) excluding the
frequency random walk. The parameters are as follows: z = 5.9 x s,
nz = 62300, v,,, = 10.6 GHz, Q = 8300, bo = -131 dB, bl = -98 dB.
= 8 x 10-l4 with bo = FkT/Po and using the microwave
8. Comparing (bo)ampli
power Po = 2 x 1oW7W (-37 dB m, as published) we find
measured at loy7 W (-40 dB m), which scale to
F =4 (6 dB).

This value includes thermal noise, the white noise of the amplifier, and the photode-
tector's shot noise.
9. Equation (5.69) gives an expression for the phase-noise transfer function
IH(j2nf ) l2 I. Fitting the phase-noise spectrum of Fig. 6.16 to the model
at the actual operating power, 2 x loe7 W (-37 dB m). Notice that the phase
noise scales with power while the fliclter phase noise does not, as explained in
Section 2.4.
12. There is a systematic inconsistency of 11 dB between our values, calculated from the
and excluding the frequency random walk (b-41 f 4), we find oscillator phase noise, and the values obtained from Fig. 6 of [78]. The inconsistency
is the same for fliclter and white noise.
13. Having detected an inconsistency, we have to decide what to trust. So, we will
worlc from the phase-noise spectrum of Fig. 6.16, discarding Fig. 6 of [78], for the
This value includes the amplifier and the photodetector.
following reasons.
10. Feeding into lH(j2n f )12 the noise parameters bo and b-1, which we have identified
Loolting at Fig. 6 of [78], at the corner point the plotted spectrum differs from the
above, plus the filter group delay tf cx Q/(nvo) and the fiber delay td, we find the
asymptotic approximation by 7 dB instead of 3 dB.
noise model shown in Fig. 6.17.
The noise spectrum is said to have been measured at -40 dB m input power. The
11. Figure 6 of [78], not shown here, reports the phase noise of the RF chain, suggesting
floor of - 139 dB rad2/Hz, found on the figure, is incompatible with the carrier
the parameters
power. At -40 dB m input power, and at room temperature, the law bo = FkT/ Po
gives a minimum floor of - 134 rad2/Hz for a noise-free amplifier.
The phase-noise spectrum of Fig. 6.16 is central in [78], while Fig. 6 is not. To
this extent, the inconsistency does not even disturb one's reading of this article.
188 Oscillator hacking 6.7 Bptoeleclronic oscillators 189

14. Introducing the responsivity p = qq/(hvl) into the expression (2.80) for the mi-
crowave power,
C)Ewaves Tidalwave photonmic oscillator (10 GHz)

(the averages are implied) we find the modulation index

The practical values for this oscillator, given in [78], are Po = 2 x W


(-37 dB m microwave power), Ro = 50 S-2 (the characteristic impedance for all
microwave circuits), p = 0.17 A/W (the photodetector responsivity), and Pl = 1.75
mW (the optical power). It follows that
offset frequency, Hz
m rti 0.3 (modulation index) Figure 6.18 Phase-noise spectrum of the Tidalwave photonic oscillator (patents 5 777 778,
6 928 091, and 7 184 45 1). Courtesy of OEwaves Inc., 2008. The interpretation, comments, and
15. The photodetector responsivity, p = 0.17 A/W and also the modulation index m z any mistaltes are those of the present author.
0.3 are unusually low. This explains the low microwave power.
16. The white and fliclter frequency noise, transformed into the Allan variance
(Table 1.4), is The phase-noise spectrum (Fig. 6.18) is fitted by the polynomial c:=-, bi f i , with

(white) , bo -138 dB 1.6~10-l4 rad2/Hz


b- 1 (not visible)
b-2 (not visible)
(fliclter) ,
b-3 -21 dB 7.9x10-~ rad2/Hz
( 2 l2~ 6.58 x 1.6 b-4 +I1 dB 1.26 x 10' rad2/Hz
.;(TI = 7h V 2 rz r = 9.3 x 10-20 r , (r. walk) .
2 x (10.6 x 109)2
We interpret the noise spectrum as follows.
Thus
1. In a delay-line oscillator with delay r we expect (see Chapter 5 ) a "clean" comb
of spectral lines at fr = l/rd, integer 1 ? 1. This structure is not present in the
spectrum of Fig. 6.18. Instead, there is a series of smaller bumps and spectral lines.
and so This indicates that the Tidalwave is not a simple delay-line oscillator and that it
contains some additional circuits that reduce or almost eliminate the spectral lines
at f, = p/rd, integer p. The remedy suggested by the literature is the dual-loop
oscillator, in which a second delay line is used to significantly reduce the largest
pealts. The references [I 11,291 provide information that is interesting yet insufficient
by far for identifying the internal details of this specific oscillator.
OEwaves Tidalwave (10 GHz OEO)
2. On the right-hand side of the spectrum, i.e., in the white-noise region, a horizontal
Apart from laboratory prototypes, OEwaves is up to the present time the only manufac- bo line gives only a basic approximation to the spectrum. This is an inherent feature
turer of OEOs. When studying oscillator noise, the Tidalwave is a challenging example of the delay-line oscillator. Discarding the pealts, we take b0 1.6 x 10-l4 rad2/Hz
because the general experience earned with traditional oscillators is now of scarce utility. (- 138 dB rad2/HZ).
190 Oscillator hacking 6,7 Opltoellectronic oscillators 191

3. The equivalent-noise figure used to estimate the microwave power must be referred 6.2 Hack as many microwave oscillators as you can, chosen in your domain of interest.
to the lowest-power point of the circuit, which is the output of the photodetector.
The reader will begin to realize that there is very little information on the web and that
The noise includes at least the intensity fluctuation of the laser, the shot noise,
the literature does not help much. Trade secrets are everywhere. Quartz and microwave
and the noise of the microwave amplifier. Let us guess that F = 6 dB. Conse-
oscillators are often based on old ideas and new technology. Many delay-line oscillators
quently, the microwave power Po = FkT/bo at the phodetector output is Po = 1 pW
are found in optics because most lasers are actually oscillators of this type. In lasers, it
(-30 dB m).
is generally difficult or impossible to separate the resonator from the amplifier.
4. One may thinlc that the transition f -3 -+f around f = 7 lcHz is the signature of
the frequency flicker of the optical fiber, cf. the frequency fluctuation of the resonator
Dear reader, thank you for getting to this page. From now you are alone in the real
in the case of the Wenzel oscillator (Section 6.4). Yet an alternative interpretation is
world. . .
possible here because f, and fL might be close to one another and fall in this region.
5. A corner frequency f, % 7 lcHz with b0 = 1.6 x 10-l4 rad2/Hz requires that b-1 =
10-lo r a d 2 / ~ (-
z 100 dB rad2/Hz). The latter value can be ascribed to the microwave
amplifier (Table 2.1) or to the optical system.
6. The delay t d is related to the Leeson frequency by fL = 1/(2nzd). Thus fL 7 kcHz
requires that rd % 23 ps, corresponding to a fiber length I = czd/n 4.7 lun, which
is a lilcely value.
-'
7. The f -+ f O transition at f 7 kHz can be ascribed either to the fact that f, % f L
or to the fliclcering of the delay. The analysis of the spectrum is not sufficient for us
to be able to say more.
8. The frequency flicker and random walk, transformed into the Allan variance
(Table 1.4), is

thus

and so

Exercises

6.1 Find on the web the phase-noise spectra s f some oscillators analyzed in this chapter.
Repeat the analysis yourself and compare the results. If possible, fit the spectra with a
power law using a set square and rule. Feel free to use Chapters 1 to 3 but refrain from
using this chapter until you have finished.
Appendix A Lap ace transforms Table A.1 Properties of Laplace transforms

Property F(s) Condition

linearity

convolution

derivative

This appendix provides a very short summary of useful properties and formulae, focusing integral
on the first-order low-pass system function (Fig. A. 1) and on the second-order resonator
system function (Fig. A.2). The reader is encouraged to refer to the numerous textboolts
time stretch
that are available, among which we prefer [66, 971.
The formulae for the Laplace transform of a function f (t) and for the inverse trans-
formation are
frequency stretch
if (I)
F(s) = L{ f(t)} = 10
00

f (t)e-"dt (Laplace transform) (A. 1) time translation f (t - a ) e-" F(s)

f (t) = L-'{ f (t)} = 1 a-jcx


~(s)e"ds (inversion formula) . (A.2)
frequency translation e-"'f ( t ) F(s - a)

The Laplace transform can be seen as an extension of the Fourier transform, on replacing
JCL)by s = a + jco, and is particularly useful when f (t) is a causal function, i.e. f (t) = 0 Table A.2 Useful Laplace transforms
for t < 0. Actual time-domain systems are always causal. That being said, the relation
between the Fourier and Laplace transforms is found as follows: Function f( 0 F(s) Condition

.F{f(t)}=IW f(t)e-jmtdt (Fouriertransform) (A.3) Dirac


-00

= 1°0 f(t1 e-jat d t


(causal signal,
f(t) = 0 for t < 0)
Dirac (delayed)

Heaviside

exponential
Table A. 1 gives properties of Laplace transforms and Table A.2 lists the transforms
of various functions. sinusoid sin wt

cosinusoid cos wt

damped sinusoid eCa' sin o t

damped cosinusoid ePatcos wt


194 Appendix A Laplace transforms Appendix A baplaee transforms 195

1.O - exp(-tlz) 1.o - cos wt exp(-tlz)


- tangent, 1 - tlz
--.
\ -' envelope, exp(-tlz)
envelope tangent, 1 - tlz
0.8
0.5

0.0

-0.5

3 - cos wt exp(-tlz)
- - envelope, exp(-tlz)
2

time, tlz

cos wt exp(-tlz)
envelope, exp(-tlz) A
tangent, 1 - tlz
envelope tangent, 1 - tlz
151 m -
10

-10

-20' time, tlz

Figure 8.11 Complex-plane pattern and impulse response of the first-order (low-pass) system Figure A.2 Complex plane pattern and impulse response of the second-order (resonator) system
+
function F(s) = l/(s l/z). Poles are represented on the complex plane as crosses. + +
function F(s) = s/(s2 2s/r m i ) . Poles and zeros are represented on the complex plane as
crosses and circles, respectively.
References [I81 S. Chang, A. G. Mann, A. N. Luiten, and D. G. Blair. Measurements of radiation pressure
effect in cryogenic sapphire dielectric resonators. P11j)s. Rev. Lett., 79(11): 2 141-2 144,
Sept. 1997. 1.6.4
[19] C. J. Chsistensen and G. L. Pearson. Spontaneous resistance fluctuations in carbon mi-
crophones and other granular resistances. Bell Sjist. Techn. J , 15(2): 197-223, Apr. 1936.
2.1.3
[20] G. Cibiel, M. Rkgis, 0.Llopis, A. Rennane, L. Bary, R. Plana, ICersalk, and \i Giordano.
Optimization of an ultra low-phase noise sapphire: SiGe HBT oscillator using nonlinear
CAD. IEEE li.arls. Ultras. Feri-oelec. and Freq. Cor~tr:,5 l(1): 3 3 4 1 , Jan. 2004. 2.5.6
[2 11 littp://m~w.vv.zaltou.coixand http://www.cmac.co~ii/.2.5.6, Table 6.1
[22] J. D. Cressler, editor. Silicon Heter.ostr.trcture Handbook. CRC, 2006. 2.5.6
Sections, szibsections, Jigtires, and tables in which a publication is mentioned are listed [23] J. D. Cressler and G. Niu. Silicorz-Germaizillr~~Heter.ojzrizction Bipolar. irl.ci1zsistor.s. Artech
irz order of occurrence. House, 2003. 2.5.6
[23] B. Cretin and J. C. Mollier. Design of a 312-order phase-locked loop for improved laser
[I] AML Communications Inc., Camarillo CA, http: //www.amlj .corn/. Table 2.2, 2.5.4,
probe resolution. IEEE Tlans. Instrtrnz. Meus., 34(4): 660-664, Dec. 1985. 6.6
Fig. 2.17
[25] S. T. Dawltins, J. J. McFerran, and A. N. Luiten. Considerations on the measurement of the
[2] P. Ashburn. SiGe Heterojunction Bipolar Transistors. Wiley, 2003. 2.5.6 stability of oscillators with frequency counters. IEEE Trans. Ultras. Feri.oelec. a11d Fi-eq.
[3] J. A. Barnes and D. W. Allan. Variances based on data with dead time between the mea- Contr:, 54(5): 9 18-925, May 2007. 1.7
surements. Technical note 1318, NIST, Boulder CO, 1990. 1.7.l
[26] R. W. P. Drever, J. L. Hall, F. \! Ibwalslti, J. Hough, G. M. Ford, A. J. Munley, and H. Ward.
[4] J. Bernamont. Fluctuations de rksistance dans un conducteur mktallique de faible volume. Laser phase and frequency stabilization using an optical resonator. Appl. Plzjis. Lett., 3 l(2):
Conzptes Rerzdz~sAcad. Sciences (Paris), 198: 1755-1758, 1934. 2.1.3
97-105, June 1983. Fig. 3.18, 3.4.3
[5] J. Bernamont. Fluctuations in the resistance of thin films. Proc. Phys. Soc. (London), 49
[27] M. M. Driscoll and R. W. Weinert. Spectral performance of sapphire dielectric resonator-
suppl.: 138-139, July 1937. 2.1.3
controlled oscillators operating in the 80 I< to 275 I< temperature range. In Pi~oc.Freq.
[6] E. D. Black. An introduction to Pound-Drever-Hall laser frequency stabilization. Anz. J Contr.01Sjinlp., San Francisco CA, May-June 1995, pp. 40 1-412. 2.5.1, 2.5.3
Phys., 69(1): 79-87, Jan. 2001. Also LIGO technical note LIGO-T980045-OOD,May 1998.
[28] B. C. Edwards, M. I. Buchwald, and R. I. Epstein. Development of the Los Alamos solid-
3.4.3
state optical refrigerator. Rev. Sci. Ii~strum.,69(5): 2050-2055, May 1998. 2.6.6
[7] H. S. Blaclc. Translation system. US Patent no. 1 686 792, issued 9 Oct. 1928 (concept of
[29] D. Eliyahu and L. Malelti. Low phase noise and spurious level in multi-loop opto-electronic
feedfonvard amplifier). 2.5.2
oscillators. In Proc. Europ. Freq. Time For.ur~zand Fi-eq. Control Sju71p. Joint Meeting,
[8] H. S. Black. Stabilized feed-back amplifiers. Electr~icalEng., 53(1): 114-120, 1934. Tampa FL, May, 2003, pp. 405410.
Reprinted in Proc. IEEE, 87(2): 379-385, Feb. 1999.2.5.2
[30] R. I. Epstein, M. I. Buchwald, B. C. Edwards, T. R. Gosnell, and C. E. Mungan. Observation
[9] R. B. Blaclunan and J. W. Tuclcey. The Measzirement of Power Spectra. Dover, 1959. 1.4.1 of a laser-induced fluorescent cooling of a solid. Nc~ture,377(6549): 500-503, Oct. 1995.
[lo] M. Born and E. Wolf. Principles of Optics. Cambridge University Press, seventh edition, 6.6
1999. 5.2.1
[3 11 J. I<. A. Everard and M. A. Page-Jones. Ultra low noise microwave oscillators with low
[I I] R. Boudot. Oscillatezirs nzicro-onde a hazite ptireti spectrale. Ph.D. thesis, Universitk de residual flicker noise. In Pi.oc. lizt. Micru~v.Theory Teclz. Sjin~p.,May 1995, pp. 693-696.
Franche Comtk, Besangon, Dec. 2006. Fig. 2.8
2.5.1
[12] R. Boudot, Y. Gruson, N. Bazin, E. Rubiola, and V. Giordano. Design and measurement of [32] W Feller. Arz Ilztroductioiz to Probability Tlzeory a i d Its Applicatioizs, vol. 2. Wiley, second
a low phase-noise X-band oscillator. Electron. Lett., 42(16): 929-93 1, Aug. 2006. 2.5.6
edition, 1971. 1.3
[I31 S. Bregni. Synclzronization of Digital Teleconznzzrizications Networlns. Wiley, 2002. 1.7.2
[33] H. T. Friis. Noise figure of radio receivers. Proc. IRE, 32: 419422, July 1944. 2.2.2
[14] 0 . E. Brigham. Tlze Fast Fourier li.ansfornz and its Applications. Prentice-Hall, 1988. 1.4.3
[34] Z. Galani, M. J. Bianchini, R. C. Waterman, Jr., R. Dibiase, R. W. Laton, and J. Bradford
[I51 J. R. Burgoon and R. L. Wilson. SC-cut quartz oscillator offers improved performance.
Cole. Analysis and design of a single-resonator GaAs FET oscillator with noise degenera-
Hewlett Packcard J : 20-29, Mar. 1981. Fig. 6.8, 6.4, Table 6.1
tion. IEEE Trnns. Micr-OM). Tlzeory Tech., 32(12): 1556-1 565, Dec. 1984. Fig. 3.20, 3.4.3,
[16] V; Candelier, P. Canzian, J. Lamboley, M. Brunet, and G. Santarelli. Ultra stable oscillators.
4.5.3
In Proc. Ezii~op.Freq. Tinze Forzim and Freq. Control Synzp. Joint Meeting, Tampa FL, May
[35] S. Galliou, F. Sthal, J.-J. Boy, and M. Mousey. Recent results on quartz crystal LD-cuts
2003, pp. 575-582. Table 6.1
operating in oscillators. In Proc. Ulti.ason. Fei+r.oelec.F14eq. Contr: Joirlt Cot$, Montreal,
[17] V. Candelier, J. Chauvin, C. Gellk, G. Marotel, M. Brunet, and R. Petit. Ultra stable Aug. 2004, pp. 475477. 6.4, Table 6.1
oscillators. In Proc. Ezirop. Freq. Time Forunz, Warszawa, Mar. 1998, pp. 345-35 1. Table
[36] E. A. Gerber and A. Ballato. Precision Freqzrerzcy Coiztr.ol (two volumes). Academic Press,
6.1
References 199

[37] L. J. Giacoletto. Study of p-n-p alloy junction transistor from dc through medium frequen- [58] J. B. Johnson. Thermal agitation of electricity in conductors. Plzj)s. Rev., 32: 97-1 09, July
cies. RCA Rev., 15: 506-562, Dec. 1954. 2.2.3 1925. 2.1.12
[38] M. L. Gorodetslty, A. A. Savchenltov, and V. S. Ilchenko. Ultimate Q of optical microsphere [59] D. ICajfez and P. Guillon, editors. Dielectric Resorzators. Noble, second edition, 1998.
resonators. Optics Lett., 21(7): 453455, Apr. 1996. 5.2.2 Previously published by Artech House, 1986, and Vector Fields, 1990. 6.6, 6.6
[39] C. A. Greenhall. Spectral ambiguity of Allan variance. IEEE Pans. Instrzlr~z.Meas., 47(3): [60] R. I<. ICarlquist. A new type of balanced-bridge controlled oscillator. IEEE Pans. Ultras.
623-627, June 1998. 1.8 Ferraoelec. and Freq. Corztr:, 47(2): 390403, Mar. 2000. Fig. 6.9, 6.4, 6.4. Table 6.1
[40] I. S. Grudinin, A. B. Matsko, A. A. Savchenltov, D. Streltalov, V. S. Ilchenko, and L. Malelti. [61] H. G. Kimball, editor. Handboolc of Selection aizd Use of Precise Freqz~erzcyand Time
Ultra high Q crystalline microcavities. Optics Cornnz., 265: 33-38, Sept. 2006. 5.2.2 Systems. ITU, 1997. 1.6
[41] A. Gruhle and C. Mahner. Low I / f noise SiGe HBTs with application to low phase noise [62] V; F. ICroupa. Theory of l/f noise - a new approach. Phj)s. Lett. A, 336: 126-132, Jan.
microwave oscillators. Electron. Lett., 33(24): 2050-2052, Nov. 1997. 2.5.6 2005. 6.3
[42] N. Gufflet, R. Bourquin, and J.-J. Boy. Isochronism defect for various doubly rotated cut [63] D. B. Leeson. A simple model of feed back oscillator noise spectrum. Pr.oc. IEEE, 54:
quartz resonators. IEEE Duns. U1tr.a~.Ferroelec. and Freq. Contr: , 49(4): 5 14-5 18, Apr. 329-330, Feb. 1966. 3.2
2002. 1.6.4, 6.4 [64] D. B. Leeson and G. F. Johnson. Short-term stability for a Doppler radar: requirements,
[43] E. A. Guillemin. Sjlizthesis of Passive Nettoorks. Wiley, 1957. 4.8 measurements, and techniques. Proc. IEEE, 54: 244, Feb. 1996 (document), 3.2.
[44] H. K. Gummel and H. C. Poon. An integral charge control model of bipolar transistors. [65] R. Leiec SiGe silences YIG oscillator phase noise. Electronic Design, Online ID no. 11902,
Bell Syst. Techn. J., 49: 827, MayIJune 1970. 2.2.3 Jan. 2006. 2.5.6
[45] D. Halford, A. E. Wainwright, and J. A. Barnes. Fliclter noise of phase in RF amplifiers: [66] W. R. LePage. Conzplex Varial?les and Laplace Parzsform for Engineers. McGraw Hill,
characterization, cause, and cure. In Proc. Fr.eq. Corztrol S'onp., Apr. 1968, pp. 340-341. 1961. Reprinted by Dover Boolts. Appendix A
Abstract only is published. 2.3.2 [67] A. N. Luiten, editor. Fr.eqztency Measzirenzent and Control. Topics in Applied Physics.
[46] A. Hati, D. Howe, D. Wallter, and F. Walls. Noise figure vs. PM noise measurements: Springer, 2001. 6.6, 6.6
a study at microwave frequencies. In Proc. Ezlrop. Freq. Tir~zeFor-irnz arzd Fr-eq. Control [68] A. B. Matslto and V. S. Ilchenko. Optical resonators with whispering-gallery modes - Part
Synzp. Joint Meeting, Tampa FL, May 2003. 2.3.2 I: Basics. J. Selected Topics Qtiarztzlnz Elec., 12(1): 3-14, Jan.-Feb. 2006. 5.2.2
[47] H. Hellwig (chair.). IEEE Starzdur.d Dejnitions of Plg)sical Qtlarztities for Fzlrzdar~zerztal [69] A. L. McWhorter. I /f noise and germanium surface properties. In R. H. Kingston, editor,
Fr-equerzcy nizd Tirne Metr.ology. IEEE Standard 1139-1 988. 1.6.1 Senzicorzd~rctorSzrrface Physics, pp. 207-228. University of Pennsylvania Press, 1957.
[48] W. R. Hewlett. A New gipe Resistance-Cupacioi Oscillator. Graduate thesis, Leland Stan- 2.1.3, 2.5.5
ford Junior University, Palo Alto CA, June 1939. 4.3.2 [70] G. I<. Montress, T. E. Parker, and M. J. Loboda. Residual phase noise of VHF, UHF, and
[49] F. N. Hooge. 1/f noise is no surface effect. Plzys. Lett. A, 29: 139-140, 1969. 2.1.3,2.5.5 microwave components. IEEE Puns. Ultras. Fen-oelec. a r ~ Freq. d C o n t ~41(5):
, 664679,
[50] C. H. Horn. A carrier suppression technique for measuring SIN and carrierlsideband Sept. 1994. 2.3.4
ratios greater than 120 dB. In Proc. Freq. Corztivl Synzp., Fort Monrnouth NJ, May 1969, [71] C. W. Nelson, A. Hati, D. A. Howe, and W. Zhou. Microwave optoelectronic oscillator
pp. 223-235. 1.9 with optical gain. In Proc. Intl. Freq. Control S'lnzp. and Ezrraop. Freq. Tinze Fortlnz Joint
[51] http://www.hparchive.co~n/Ma~i~~als/. 6.4 Meeting, Geneva, May-June 2007, pp. 1014-1019. 2.6.2
[52] V. S. Ilchenko and A. B. Matslto. Optical resonators with whispering-gallery modes - Part [72] H. Nyquist. Thermal agitation of electric charge in conductors. Plzys. Rev., 32: 110-1 13,
11: Applications. J. Selected Topics Qtlur~tzrrnElec., 12(1): 15-32, Jan.-Feb. 2006. 5.2.2 July 1928. 2.1.1
[53] V. S. Ilchenlto, X. S. Yao, and L. Malelti. Pigtailing the high-Q n~icrospherecavity: a [73] http://www.oscilloquartz.coni/. Table 6.1
simple fiber coupler for optical whispering-gallery nlodes. Optics Lett., 24(11): 723-725, [74] A. Papoulis. Pr+obability,Random Variables and Stochastic Processes. McGraw Hill, third
June 1999. 5.2.2 edition, 1992. 1.3
[54] E. N. Ivanov, J. G. Hartnett, and M. E. Tobar, Cryogenic microwave amplifiers for precision [75] D. B. Percival and A. T. Walden. Spectral Analysis for Phj~sicalApplications. Cambridge
measurements. IEEE Trans. Ultras. Fer*roelec.and Freq. Contr: , 47(6): 1273-1274, Nov. University Press, 1998. 1.4.1, 6.1.2
2000. Fig. 2.9, 2.3.3 [76] N. Pothecary. Feedforfi~vurdLinear Power Anzp1iJier.s. Astech House, 1999. 2.5.2
(551 E. N. Ivanov, M. E. Tobar, and R. A. Woode. Application of interferometric signal process- [77] R. V. Pound. Electronic frequency stabilization of microwave oscillators. Rev. Sci.
ing to phase-noise reduction in lnicrowave oscillators. IEEE Purzs, Adicrow. Tlzeory Tech., Instriinz., 17(11): 490-505, Nov. 1946. Fig. 3.17, 3.4.3,4.5.3
46(10): 1537-1545, Oct. 1998. 2.5.3 [78] S. Romisch, J. Kitching, E. F. Piltal, L. Hollberg, and F. L. Walls. Performance evaluation
[56] G. M. Jerkins and D. G. Watts. Spectr*alArzalysis a r ~ dits Ayylicatiorzs. Holden Day, 1968. of an optoelectronic oscillator. IEEE Trans. Ultras. Ferroelec. arzd Freq. Contr:, 47(5):
1.4.1, 6.1.2 1159-1 165, Sept. 2000. 6.7, Fig. 6.16
[57] J. B. Johnson. The Schottky effect in low frequency circuits. Plzys. Rev., 26: 71-85, July [79] E. Rubiola. The lneasurement of AM noise of oscillators. Website arXiv.org, document
1925. 2.1.3 arXiv:physics/05 12082, Dec. 2005. 1.6.4, 3.1.1
200 References References 201

[80] E. Rubiola. On the ineasureinent of frequency and of its sample variance with high- [99] I<. K. Thladhar, G. Jenni, and J. Aubry. Improved BVA resonator - oscillator perforn~ances
resolution counters. Rev. Sci. I11strtm1.,76(5), May 2005. Websites arxiv.org and rubiola.org, and fi-equencyjumps. In Proc. Ez~rop.Frey. Tinze For-t~nz,Neuchiitel, Mar. 1997, pp. 273-
document arXiv:physics/0411227, Dec. 2004. 1.7 280. Table 6.1
[8 11 E. Rubiola and R. Boudot. The effect ofAM noise on correlation phase noise nzeasurenzents. [loo] A. van der Ziel. Noise irz Solid State Devices and Circuits. Wiley, 1986. 2.2.3, 2.2.3
Website arxiv.org, document arXiv:physics/0609147, Sept. 2006. 1.9.2 [ l o l l M. E. Van Valltenburg. Modern Network Synthesis. Wiley, 1960. 4.8
[82] E. Rubiola and V. Giordano. Advanced interferometric phase and amplitude noise mea- [I021 J. Vanier and C. Audoin. The Quantunz Physics of Atonzic Frequency Standards. Adam
surements. Rev. Sci. Instrw~z.,73(6): 2445-2457, June 2002. Website arxiv.org, document Hilger, 1989. 1.6
arXiv:physics/0503015vl.2.5.3 [103] J. R. Vig (chair.). IEEE Standard DeJinitions of Phjisical Quantities for Fundamental
[83] E. Rubiola and V. Giordano. On the l/f frequency noise in ultra-stable quartz oscillators. Frequency and Tinze Metrology - Rarzdonz Instabilities (IEEE Standards 1139-1 999).
IEEE Puns. Ultras. Ferroelec. arzd Fr-eq. Contr:, 54(1): 15-22, Jan. 2007. Fig. 2.7, Fig. IEEE, 1999. 1.6, 1.6.1, 1.6.1
3.1 1, Fig. 6.2, 6.3, Table 6.1 [104] F. L. Walls. The quest to understand and reduce l / f noise in amplifiers and BAW quartz
[84] E. Rubiola, J. Groslambert, M. Brunet, and V. Giordano. Fliclter noise measurement of HF oscillators. In Proc. Europ. Fr.eq. Time Fortrr~z,Besan~on,Mar. 1995, pp. 227-243. 6.3
quartz resonators. IEEE Trar~s.Ultr*as.Fer-roelec. and Fkeq. Contr:, 47(2): 361-368, Mar. [105] F. L. Walls, E. S. Ferre-Piltal, and S. R. Jefferts. Origin of l/f PM and AM noise in
2000. 6.3 bipolar junction transistor amplifiers. IEEE Pans. Ultr*as. Ferroelec. and Freq. Corztr:,
[85] E. Rubiola, Y. Gruson, and V: Giordano. On the fliclter noise of circulators for ultra-stable 44(2): 326-334, Mar. 1997. 2.3.2
oscillators. IEEE TI.ar?s. Ultreas. Fer.roelec. and Freq. Corztr: , 5 l(8): 957-963, Aug. 2004. [106] L. Weinberg. Network Analjisis and Sjintlzesis. McGraw Hill, 1962. 4.8
Fig. 3.17, Fig. 3.20 [I071 http://www.wenzel.coni. Table 6.1
[86] E. Rubiola and F. Lardet-Vieudrin. Low fliclter-noise amplifier for 50 S2 sources. Rev,Sci. [lo81 R. A. Woode, M. E. Tobar, E. N. Ivanov, and D. Blair. An ultra low noise microwave
Instr4tm~., 75(5): 1323-1 326, May 2004. Free preprint available on the website aixiv.org, oscillator based on high Q liquid nitrogen cooled sapphire resonator. IEEE Pans. Ultras.
document arXiv:plzysics/050301 2 ~ 1Marclz , 2005. 2.3.4 Fe~roelec.and Freq. Contr:, 43(5): 936-941, Sept. 1996. Fig. 6.15, 6.6
[87] E. Rubiola, E. Salilt, N. YLI,and L. Malelti. Phase noise measurement of low-power signals. [I091 X. S. Yao, L. Davis, and L. Malelti. Coupled optoelectronic oscillators for generating both
In Proc. Eta-up. Fr-eq. Erne For.tlnz and Freq. Control Syrnp. Joint Meeting, Montreal, RF signal and optical pulses. J Lightwave Technol., 18(1): 73-78, Jan. 2000. 4.4.1
Aug. 2004, pp. 292-297. Fig. 2.6 [I101 X. S. Yao and L. Malelu. Optoelectronic microwave oscillator. J Opt. Soc. Am. B - Opt.
[88] E. Rubiola, E. Salilt, N. Yu, and L. Malelti. Fliclter noise in high-speed p-i-n photodiodes, Phj~s.,13(8): 1725-1735, Aug. 1996. 5.1, 5.8, 6.7
IEEE Pans. &ficr*o~ t ~ . Tl~eoqiTech., 54(2): 8 16-820, Feb. 2006. Preprint available on [ I l l ] X. S. YBo and L. Maleli. Multiloop optoelectronic oscillator. J Quantuln Electron., 36(1):
arxiv.org, document arXiv:physics/0503022v 1, March 2005. 2.6.2 79-84, Jan. 2000. 6.6
[89] J. Rutman. Characterization of phase and frequency instabilities in precision frequency
sources: fifteen years of progress. Proc. IEEE, 66(9): 1048-1075, Sept. 1978. 1.6
[90] B. E. A. Saleh and M. C. Teich. Ftrr~dume~ztals of Photolzics. Wiley, 1991. 5.2.1
[91] A. S. Sedra and K. C. Smith. Microelectr.onic Circuits. Oxford University Press, fifth
edition, 2004. 3.1
[92] H. Seidel. A microwave feedfoiward experiinent. Bell Sjjst. Techn. J , 50(9): 2879-2916,
Nov. 1971.
[93] W. Shieh, X. S. Yao, L. Malelti, and G. Lutes. Phase-noise characterization of optoelectronic
conzponents by carrier suppression techniques. In Proc. Optical Fiber Conzm. Con$, San
Josk CA, May 1998, pp. 263-264. 2.6.2
[94] A. E. Siegman. Laser-s. University Science Boolts, 1986. 5.2.1
[95] J. Silala, S. Hashiguclzi, M. Ohki, and M. Tacano. Some considerations for the construction
of low-noise amplifiers in very low frequency region. In J. Silula and M. Levinsthein, edi-
tors, Advarzced Experin~entalMethods ir~Noise Researclz orz Narzoscale Electr.orzicDevices,
pp. 237-244. ICluwer, 2004. 2.3.4
[96] J. J. Snyder. Algorithm for fast digital analysis of interference fringes. Appl. Opt., 19(4):
1223-1225, Apr. 1980. 1.7.2
[97] M. R. Spiegel. Schclzrrn 5 Otltline of Laplace Pansfornls. McGraw Hill, 1965. Appendix A
[98] P. Sulzer. High stability bridge balancing oscillator. Proc. IRE, 43(6): 701-707, June 1955.
Fig. 3.19, 3.4.3
ndex noise (cont.)
shot, 37,41,64
relaxation time, 72, 91
repeatability, 11
thermal, 35-41 reproducibility, 11
noise figure, 38,43,46,49, 54, 57, 58, 153 resonator, 88-96
noise temperature, 37 Fabry-PCrot, 128-129, 143
free decay of, 88-9 1, 103-104
optoelectronic oscillatol; 55, 62-63, 146-147 mode selection, 133-138, 140-143
examples, 185-190 phase response of, 101-1 10
oscillator whispering-gallery, 129-1 30
Barlchausen condition, 67, 93, 98, 125, 126 resonator noise, 79-8 1
delay-line, 82-83, 125-127
frequency noise in, 73,83, 84, 116, 145 spectrum, 13-1 9
frequency pulling of, 70-72,114, 137 estimation of, 151-152
ADEV, see variance, Allan frequency noise frequency-loclted, 83-87 interpretation of, 75-82, 153-1 54,
amplifier flicltel; 23, 30, 31, 76, 78, 80, 151 loop gain of, 98, 100, 101, 127 172-175
buffer, 76-79,154156 power-law, see spectrum, power-law Pound stabilization of, 85-87 one-sided, 14, 15
cascaded, 40,49-52, 154-156 random walk, 23,30, 3 1, 151 starting, 69 power law for, 23,24, 30, 3 1, 151
feedforward, 53-55 white, 23, 30, 31, 76, 78, 80, 151 relation to variances, 29, 30, 3 1
noise-corrected, 55-57, 78, 79, 153 frequency response, see transfer function phase noise, 3-5,6,22-23 stationary ergodic signals in, 15
parallel, 57-60 frequency-locked loop (FLL), see oscillator, of cascaded amplifier, 49-52 two-sided, 15
push-pull, 58 frequency loclted corner frequency in, 46-48,75-76,78-79, 153 spectrum analyzer, 16-1 9, 30
SiGe, 46,61 Fresnel vector, 1 flicltel; 23, 30, 31,43-48,49, 52, 61, 76, 78, 80, FFT, 18,32
sustaining, 67, 75, 8 1-82, 156 Friis formula, 40, 49 151 stationary random process, 11
transposed gain, 52 frequency synthesis of, 8-9 stochastic process, see random process
amplitude noise, 3-5, 6, 24,44, 100 gain control, 69-70, 100-101, 144-145 measurement of, 30-33 system function, see transfer function
analytic signal, 2 photodetector, 63
AVAR, see variance, Allan impulse response, 19, 101, 105, 106, 108, 194, power law, see spectrum, power-law transfer function, 19,2 1
average, 10 195 white, 23,30,3 1,42-43,49, 76, 78, 80, 151 oscillator, 97-100, 125-127, 130, 134
time, 10 phase time, 7 phase-noise, 74, 88, 101, 105, 106, 109, 111-1 12,
weighted, 10 =wf 1322
Laplace transform, 20-22, 192
phase-locked loop (PLL), 3 1-33 113, 117, 123
phasor, 1 resonator, 67,93-94, 129
Barlchausen condition, see oscillator, Barlchausen low-pass, 194 time-varying, 2, 106, 108 roots of, 92, 93, 98-100, 110, 114, 119, 124, 126,
condition resonator, 92-96, 1 0 4 110, 195 photonic oscillator, see optoelectronic oscillator 131,133, 134-137, 140-142
lases, 87, 143-145 power spectral density (PSD), see spectrum symmetry of, 95-96
clock signal, 3 Leeson effect, 72-75,78-79,82,85, 11 1-1 12, 142, pseudofrequency, see free decay
complex envelope, 2 154156 variance, 25-30
cyclostationary process, 11, 13 low-pass process, 2, 6, 105, 121 quality factor Q, 71, 82, 84, 88-9 1, 95, 101, 142, Allan, 25,27, 116, 145-146, 153-154
153,155-156 classical, 25
damped oscillation, see free-decay pseudofrequency MDEY see variance, modified Allan quartz oscillator, 68, 72, 79, 115, 154-175 modified Allan, 27-29
detuning, 106-1 10, 113-1 14, 137 mean, 10 examples, 156-172, 174 relation to spectra, 29, 30, 3 1
deviation, see variance microwave oscillator, 55, 78-79, 115 square of deviation, 25
dissonance, 94, 123, 134 examples, 175-1 85 random process, 10 transfer function, 26,28
mixer (double-balanced), 3 1 measurement, 13 wavelet, 26,28
ergodicity, 11 MVAR, see variance, modified Allan
excess noise, see noise, fliclter
expectation, 11 near-dc process, 6, 11
noise
Fourier transform, 15-22, 192 additive, 36,42,49
free decay pseudofrequency, 8 , 9 1, 105 environment (effect of), 36,48,5 1
frequency, 1 flicker, 36, 37
angular, 1, 7 frequency, 6,24
carrier, 6, 8 in-phase, 4
close to dc, 6 1/f , see noise, fliclter
fractional, 8 parametric, 36,43,45
Leeson, see Leeson effect photodetector, 4 1
natural, 8, 67, 83, 89, 98, 104 quadrature, 4

You might also like