Dissertation Erb Final
Dissertation Erb Final
Dissertation Erb Final
Stefan Erb
Jitter Analysis Methods for the Design and
Test of High-Speed Serial Links
Distribution Tail Fitting Based on Gaussian
Design Quantile
of EmbeddedNormalizationCMOS
A/D-Converters
for Communication Systems
Dipl.-Ing. Stefan Erb
————————————–
GrazGraz
University ofTechnology
University of Technology
ife
institut für elektronik
Institute of Electronics
Institute of Electronics
Supervisor: .........................................................................
Univ.-Prof. Dipl.-Ing. Dr. techn. Wolfgang Pribyl
Supervisor: .........................................................................
Univ.-Prof. Dipl.-Ing. Dr. techn. Wolfgang Pribyl
Co-Supervisor: .........................................................................
Univ.-Prof. Dipl.-Ing. Dr. techn. Ernst Stadlober
Co-Supervisor: .........................................................................
Univ.-Prof. Dr. Willy Sansen
Graz, May 2011
Statutory Declaration
I declare that I have authored this thesis independently, that I have not used other than the declared
sources / resources, and that I have explicitly marked all material which has been quoted either
literally or by content from the used sources.
....................... .............................
Date Signature
V
Abstract
In this thesis analysis methods for serial high-speed interfaces are presented to investigate charac-
teristics and impact of timing uncertainty or jitter. Such methods are widely used for estimating the
bit error rate (BER) of serial links, and for signal integrity measurements and compliance testing.
First, an algorithm is developed which determines the Gaussian tail behavior of measured jit-
ter distributions and separates them into random and deterministic components using an efficient
optimization scheme. The resulting timing budget allows to accurately quantify the total jitter of
clock signals and phase locked loop (PLL) systems. The fast analysis technique allows for an
implementation as embedded system and thus, supports a broad variety of design-for-test (DFT)
applications such as production testing, on-chip diagnostics and on-line monitoring. One chapter
of the thesis is dedicated to these hardware design aspects.
The underlying mathematical principle is based on the Gaussian quantile normalization, which
has already been discussed and utilized in previous approaches. A detailed and thorough perfor-
mance comparison is thus carried out to highlight their different properties. Further, it is demon-
strated how the proposed method can easily be generalized for use with arbitrary non-Gaussian
tails, as is the case for optical high-speed interconnects.
In a first case study, a fast system level model of a serial high-speed PLL is developed. The
system has already been realized as test structure and thus, allows for a direct comparison with
measurements. These include the analysis of closed loop phase noise, jitter transfer characteristics
as well as the jitter tolerance behavior. Especially the latter two also comprise the use of the
developed method. A second case study realizes a BER tester with a 3Gb/s serial high-speed
interface on an FPGA board. It demonstrates the practical aspects of jitter measurement, diagnosis
and optimization.
VII
Kurzfassung
In dieser Arbeit werden Methoden zur Analyse von Zeit-Jitter vorgestellt, die vor allem in digi-
talen seriellen Schnittstellen Anwendung finden. Solche Methoden erlauben die Abschätzung der
Bitfehlerrate (BER) und werden deshalb häufig für die Beschreibung der Signal-Qualität und zum
Nachweis der Einhaltung von Standards eingesetzt.
Im Rahmen der Arbeit wird zunächst eine Methode entwickelt, welche das Gauß-Verhalten an
den beiden Enden einer gemessenen Jitter-Verteilung feststellt und die Verteilung daraufhin ex-
trapoliert. Die Methode zeichnet sich vor allem durch ihre Genauigkeit und Effizienz aus. Sie ist
daher besonders für die quantitative Analyse der Qualität von Taktsignalen und Phasenregelschlei-
fen (PLLs) geeignet. Aufgrund des einfachen Algorithmus kann sie auch sehr gut zusammen mit
Teststrukturen, Diagnosewerkzeugen oder für die Echtzeit-Überwachung solcher Systeme einge-
setzt werden.
Das zugrundeliegende mathematische Prinzip basiert auf der Quantil-Normalisierung von Gauß-
Verteilungen, welche bereits in ähnlichen Ansätzen verwendet wurde. Deshalb umfasst die Arbeit
auch einen detaillierten und umfangreichen Vergleich verschiedener Methoden, der die jeweiligen
Eigenschaften aufzeigt. Ein weiterer Abschnitt ergänzt die vorgestellte Methode um ein generali-
siertes Prinzip, mit dem sie auf einfache Weise auf beliebige nicht-Gauß Verteilungen angewandt
werden kann.
In einer ersten Fallstudie wird ein sehr schnelles Modell für eine PLL erstellt. Dieses Sys-
tem existiert bereits als Teststruktur, und ermöglicht daher einen direkten Vergleich der erhalte-
nen Simulationsergebnisse mit Messdaten. Der Vergleich umfasst eine Rauschanalyse der PLL,
die Simulation der Jitter-Transfer-Funktion sowie des Jittertoleranz-Verhaltens. Die Simulation
der letzten beiden Eigenschaften erfolgt dabei mit Hilfe der neuen Analysemethode. Eine zweite
Fallstudie realisiert einen BER-Tester zusammen mit einer 3Gb/s S-ATA Schnittstelle auf einem
FPGA Board. Sie dient vor allem der Darstellung praktischer Aspekte bei der Diagnose und Opti-
mierung von Jitter.
IX
Danksagung
Ich möchte mich bei allen bedanken die mich in den letzten dreieinhalb Jahren bei der Umsetzung
dieser Dissertation unterstützt haben.
Allen voran, mein Betreuer und Mentor, Prof. Wolfgang Pribyl, der mir bei meinem Forschungs-
thema großen Freiraum gelassen, und mich in meinen Interessen und Anliegen sehr stark unter-
stützt hat. Diese Rahmenbedingungen haben das erfolgreiche Abschließen der Dissertation erst
möglich gemacht und mitunter ganz wesentlich zur Qualität der vorliegenden Arbeit beigetragen.
Ich empfinde den erfolgreichen Abschluss daher als großes Geschenk, das uns hoffentlich noch
viele Jahre über diese Arbeit hinaus verbindet.
Ein weiterer großer Dank gebührt meinem Zweitbegutachter, Prof. Ernst Stadlober, der mir in
zahlreichen Diskussionen die wesentlichen Aspekte der Extremwerttheorie für mein Thema be-
reitwillig nähergebracht hat. Ausgehend von der Modellierung elektronischer Systeme hat meine
Arbeit dadurch einen Streifzug durch die Welt der Statistik unternehmen können, den ich für mich
persönlich als sehr bereichernd empfinde.
Für die fachliche Unterstützung danke ich vor allem den Ingenieuren des ‘Analog Design Sup-
port’ Teams der Infineon Villach, die mir bei der Erstellung des Modells für serielle high-speed
Schnittstellen geholfen und die Messdaten der Teststruktur zur Verfügung gestellt haben. Reinhard
Steiner hat mir immer wieder gern Auskunft erteilt sowie den Zugang zu den Hochleistungsrech-
nern über die gesamte Dauer der Dissertation ermöglicht.
Nicht zuletzt verdanke ich auch meinen Kollegen aus dem Dissertantenlabor zahlreiche nütz-
liche Hinweise und Anregungen. Allen voran Christoph Böhm, der mir immer wieder gerne und
manchmal auch sehr kurzfristig meine Publikationen verbessert hat.
Ich möchte an dieser Stelle auch noch Siegfried Rainer und Jan Ranglack erwähnen, für deren
Freundschaft ich sehr dankbar bin. Danke William Robinson, für die Englisch-Korrekturen.
S TEFAN E RB
G RAZ , M AI 2011
XI
The real voyage of discovery consists not in seeking new landscapes
but in having new eyes.
M ARCEL P ROUST
K URT L EWIN
S EPP H ERBERGER
Contents
List of Figures v
List of Tables ix
List of Abbreviations xi
Nomenclature xiii
1. Introduction 1
1.1. Motivation and Problem Domain . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2. Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3. Thesis Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3. Mathematical Background 17
3.1. Quantile Normalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1.1. Quantile Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1.2. Gaussian Quantile Normalization . . . . . . . . . . . . . . . . . . . . . 18
3.2. Performance Analysis of Algorithms . . . . . . . . . . . . . . . . . . . . . . . . 21
3.2.1. Performance Metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.2.2. Test Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
I
4.3.2. Improvement of Convergence Behavior . . . . . . . . . . . . . . . . . . 45
4.3.3. Performance Analysis with Optimized Parameters . . . . . . . . . . . . 50
4.4. Performance Optimization with Q-Domain Threshold . . . . . . . . . . . . . . . 51
4.4.1. Minimum Q-Domain Threshold Definition . . . . . . . . . . . . . . . . 51
4.4.2. Parameter Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.4.3. Performance Analysis with Optimized Parameters . . . . . . . . . . . . 57
4.5. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
II
8.2.2. Jitter Transfer Function . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
8.3. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
10. An FPGA based Diagnostic Tool for Jitter Measurement and Optimization 145
10.1. Measurement Principle and Implementation . . . . . . . . . . . . . . . . . . . . 145
10.1.1. Diagnostic Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
10.1.2. Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
10.2. Jitter Measurements and Optimization . . . . . . . . . . . . . . . . . . . . . . . 148
10.3. Analysis of Extrapolation Error . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
10.4. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
Bibliography 169
III
List of Figures
3.1. Example for an empirical distribution function and corresponding quantile plot. . 18
3.2. Q-normalization principle demonstrated with a bathtub function. . . . . . . . . . 20
3.3. Simple optimization scheme for Gaussian tail fitting based on Q-normalization. . 20
3.4. Definition of interquartile range for the Gaussian distribution. . . . . . . . . . . . 22
3.5. Test distribution types, constructed with parameters σRJ and ADJ . . . . . . . . 23
V
4.22. ∆Tt versus Pt,min parameter surfaces to investigate Emed , IQR, EL and κ of three
different optimization criteria. . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.23. ∆Pt versus Pt,min parameter surfaces to investigate Emed , IQR, EL and κ of
three different optimization criteria. . . . . . . . . . . . . . . . . . . . . . . . . 48
4.24. ∆Pt versus Pt,min parameter surfaces to investigate EL of three different opti-
mization criteria at a different test distribution shape. . . . . . . . . . . . . . . . 49
4.25. Emed and EL of optimized ĉ1.2 criterion over varying σRJ /ADJ,uni and N . . . . 50
4.26. Emed and EL of optimized ĉ1.2 criterion over varying σRJ /ADJ and DJ type. . . 51
4.27. Qmin threshold parameter definition. . . . . . . . . . . . . . . . . . . . . . . . . 52
4.28. Flow graphs of algorithms based on minimum Qth,min and constant Qth,c thresh-
old in Q-domain. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.29. Emed , EL and κ of four different algorithmic principles over varying σRJ /ADJ . . 54
4.30. Emed , EL and κ of four different algorithmic principles over varying Qmin . . . . 56
4.31. Emed , EL and κ of Qth,c and σ̂err algorithm over varying Qmin and DJ shape. . . 56
4.32. Emed and EL of Qth,c and Qmin =−1.2 over varying σRJ /ADJ,uni and N . . . . 57
4.33. Emed and EL of Qth,c and Qmin =−1.2 over varying σRJ /ADJ,uni and DJ type. . 57
VI
6.13. Comparison of sQN, QN, QP2, QP3, QP4 methods at constant N =106 . . . . . . 101
6.14. Comparison of sQN, QN, QP2, QP3, QP4 methods at constant N =107 . . . . . . 102
6.15. Comparison of sQN, QN, QP2, QP3, QP4 methods at constant N =108 . . . . . . 103
6.16. EL of sQN, QN, QP2, QP3, QP4 methods over varying number of bins R. . . . . 104
6.17. Design example: EL of sQN and QN methods over number of counters C. . . . . 107
7.1. Eye diagram with timing jitter and amplitude noise PDFs. . . . . . . . . . . . . . 110
7.2. Generalized optimization scheme. . . . . . . . . . . . . . . . . . . . . . . . . . 110
7.3. Special GGD shapes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
7.4. GGD random generator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
7.5. T̂ surface for two-dimensional minimum search with initial search grid. . . . . . 115
7.6. Emed , EL and Emed,α over varying test distribution shape. . . . . . . . . . . . . 116
7.7. Emed , EL and Emed,α over varying sample size N . . . . . . . . . . . . . . . . . 117
7.8. Emed , EL and Emed,α over varying number of bins R. . . . . . . . . . . . . . . 117
7.9. Performance comparison using EL over varying jitter ratio σRJ /ADJ . . . . . . . 118
VII
10.15.TJpp medians of fitted tails, exact measurements and expected worst case error
over varying cable length. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
VIII
List of Tables
6.1. Default parameter configuration of ∆Pt for polynomial tail fitting methods. . . . 92
6.2. Emed , IQR and EL coefficients for QN method, equation (6.6). . . . . . . . . . 105
6.3. Emed , IQR and EL coefficients for QN method with included DNL error, equa-
tion (5.38). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
IX
List of Abbreviations
XI
XII
Nomenclature
XIII
p Probability
Pt,min Minimum probability threshold for tail fitting, see figure 4.21
q Quantile
Qmin Minimum Q-domain threshold for tail fitting, see figure 4.27
Qth,c Algorithmic scenario for tail fitting based on Qmin threshold, see section 4.4.2
R Number of bins per unit interval
s Slope of linear function
T Bit period, if normalized to the unit interval T =1 UI
T (fSJ ) Jitter transfer function
teye Eye opening
tL,R Left/Right timing budget of jitter, see figure 2.10
x Timing jitter amplitude
XIV
1. Introduction
A brief overview to the topic of jitter and bit error rate (BER) analysis in high-speed communica-
tions is given. The problem domain is presented and an outline to the overall document is given,
together with a list of scientific contributions which have been elaborated throughout the work.
TX Data RX Data
TX−PLL RX−PLL
Serial high-speed interfaces usually transmit data without a dedicated clock signal. A special
encoding guarantees for sufficient bit transitions inside the data stream in order to correctly syn-
chronize the receiver clock with the input data. The RX-PLL as a synchronization circuit has to
cope with input jitter and provide a certain robustness against timing variations. If jitter exceeds a
critical amplitude, the PLL will not be able to track timing variations correctly, and hence, lead to
erroneous signal recovery or misinterpretation of the received data.
Serial interface standards, such as S-ATA [118], specify stringent requirements on the error
probability, or bit error rate (BER) of a recovered data stream. The BER directly reflects the influ-
ence of timing jitter on system performance, and is thus the best suited figure of merit to indicate
1
1. I NTRODUCTION
the quality of a digital communication interface. In fact, one likes to describe a relation between
BER and measured jitter to thoroughly investigate its influences and to afford identification of
possible root causes.
Since jitter is basically a random process, its analysis involves the use of statistical analysis
methods. These are applied to a set of collected jitter samples and try to accurately determine
the impact of timing jitter on the investigated system. In this thesis such jitter analysis methods
are developed, analyzed, compared and applied to practical simulations and measurement cases.
The subsequent section describes the key contributions and results which have been achieved
throughout this work. It is followed by a brief overview of the overall document.
1.2. Contributions
Several key contributions and results of this work extend the state of knowledge, and are summa-
rized below. A more detailed description of the key results can also be found in the conclusions
section at the end of the thesis, together with a list of own publications [C1-C9].
First, a novel method (here denoted as scaled Q-normalization, sQN) for jitter and BER anal-
ysis is developed. It is based on the Gaussian quantile normalization principle, where the three
parameters amplitude, mean and standard deviation of Gaussian model functions are identified
and fitted into the tails of a jitter distribution. This allows them to be extrapolated down to any
desired probability level. The method is realized with a flexible and efficient optimization scheme,
and allows for fast tail fitting combined with accurate extrapolation results. Its performance is
analyzed with respect to the extrapolation error, which is shown to highly depend on the sample
size and the shape of test distributions. From the basic concept, two algorithmic approaches with
conservative fitting parameters are derived and optimized, in order to improve the error behavior
with respect to accuracy and outlier suppression. For typical test distributions (uniform combined
with Gaussian) and a number of jitter samples N =106 (tail extrapolations range over six orders
of magnitude, down to the 10−12 level), the estimated jitter budget has an error bias <2% and an
overall error <3% in more than 97.5% of the cases. The method is partly published in [C1,C8,C9].
Another contribution of this thesis focuses on hardware design aspects, to utilize the proposed
sQN method together with real jitter measurement systems. Their limited precision causes quan-
tization effects and introduces additional extrapolation error which must be considered and dealt
with. First, requirements with respect to minimum tail amplitude and time resolution of measured
distributions are investigated. Corresponding equations are derived to guarantee these require-
ments. Then, the combined influence of limited sample size and time resolution on the accuracy
of the sQN method is described in terms of empirical relations that quantify error bias, statistical
spread and the combination of both. These relations aid a system designer in finding an optimum
performance trade-off between desired accuracy of analysis and hardware expense. Finally, also
the effect of process variations in measurement systems is added to these empirical relations, in
order to quantify their influence. Two typical design examples act as design guidelines to real-
ize hardware jitter measurements with a certain target accuracy, when using the proposed sQN
method. The described contributions are also published in [C4,C8].
Further, a comprehensive performance comparison of different jitter analysis methods based on
the quantile normalization principle is provided. Therefore, the sQN method is compared with
the known conventional quantile normalization (QN) method, as well as higher order polynomial
methods (QP2, QP3, QP4). The idea is to provide a detailed performance reference for compar-
ison with future jitter analysis methods. As a fundamental result, in simulators the QN method
highlights the same beneficial property of a strictly positive error bias as the sQN method. This is
a clear advantage compared to higher order polynomial methods (QP2, QP3, QP4), which achieve
acceptable accuracy only for certain test distributions. A comprehensive comparison with the sQN
2
1.2. C ONTRIBUTIONS
method is carried out, and clearly shows that sQN achieves the best performance. However, this is
at the cost of a larger computational demand. Although the QN method is less accurate, it offers a
significant speed-up (≈ 35 times compared to sQN) for tail fitting. This complementary property
demands an additional error analysis with QN when used together with hardware measurements.
This allows a system designer to choose between the better suited algorithmic alternative. Equiv-
alent to the sQN method, an error analysis is thus carried out for the QN method, which quantifies
its extrapolation error in terms of the two key parameters sample size and time resolution, as re-
quired for hardware jitter measurements. In addition, the previously described design examples
for hardware measurements are extended for use with the conventional QN method. Therefore,
the design equations for the sQN method can simply be reused. Obtained results highlight that
also for hardware measurements the QN method is generally less accurate than sQN, but is also
less affected by differential non-linearity error, as caused by process variations. These results are
partly published in [C3,C8].
A dedicated part of the thesis also focuses on the generalization of the sQN principle for use with
arbitrary non-Gaussian tails. Such tails may for example appear with amplitude distributions of
high-speed optical links. As a function class, the generalized Gaussian distribution (GGD) is used
for tail fitting. The presented principle is fully consistent with the existing jitter decomposition
model and thus, forms a logical extension. In simulators, it is able to identify the exponential
tail characteristic of a distribution and to extrapolate tails with acceptable accuracy. However, the
estimated tail characteristic suffers from large errors if tails decay very fast or in hardware systems
with very limited time resolution. Further, with its large computational demand the primary use
cases are simulations. The generalized principle is also presented in [C5].
Several case studies describe typical application fields for the proposed sQN method. First, a
fast behavioral model for charge-pump PLLs is implemented, which is based on an exact solution
for the 2nd order loop filter, and includes a parasitic gain regulator pole as well as an oscillator
noise model [C2]. It is able to simulate approximately 106 bit periods within one minute on
an Intel 2.2GHz laptop and thus, allows for in-depth system exploration as well as statistical
jitter analysis. Jitter transfer functions and phase noise spectra of the modeled PLL are compared
with measurements from an existing hardware structure and show excellent agreement. The sQN
method is here used to simulate and verify the measured jitter transfer functions. Therefore it
extracts the deterministic jitter component from collected distributions. For additional comparison,
also a spectral analysis method is used.
As a second application, the sQN method is used for identifying jitter tolerance curves of the
previous high-speed PLL model. Therefore, external jitter is injected to the PLL and adjusted
until a desired error probability is obtained. In order to solve this inverse problem, an adaptive
search algorithm is developed, which highly reduces the number of required jitter samples [C6].
Comparative results show, that the included sample size adaptation makes the recursive search 2-3
times faster. Results also highlight, that the proposed algorithm can be used for both simulations
and hardware measurements.
In a final case study, a practical jitter measurement and analysis system for the diagnosis and
optimization of transmission lines, PLLs and transceiver structures is developed [C7]. The target
architecture is a high-speed FPGA, and as an example, various jitter measurements are carried out
with RG-58 coaxial cables as well as a standard 1m S-ATA cable. Optimizations are performed
with the FPGA internal transceiver settings and equalizer structures, which allow to reduce the to-
tal jitter of a 5m test cable significantly (up to 28%). In concluding analyses, also the extrapolation
error of sQN and QN fitting methods is investigated and compared against theoretical worst case
errors from previous numerical analyses. In this context the different DJ shapes are experimentally
confirmed to be well suited for estimating the error of pure PLL jitter, ISI, and ISI plus additional
noise affected channels.
3
1. I NTRODUCTION
4
2. Fundamentals of Jitter, PLLs and BER
Analysis
In this chapter an introduction to the basics of jitter and PLLs in high-speed serial links is given.
After a brief overview to the sources of timing jitter in communication systems, different types
and definitions for clock jitter and phase noise are discussed. The fundamental principle of PLLs
for high-speed data transmission is explained, followed by a comprehensive overview of the state-
of-the-art of jitter analysis methods. The main focus is on analysis techniques that relate jitter with
the bit error rate (BER), as required for testing high-speed serial links.
TX Channel RX
Clean
Data
EQ CDR
ISI
Reflection
Jittery Crosstalk accumulated
Tx−Clk jitter
ideal clock synthesizer inside the transmitter structure. Depending on the quality of the channel,
inter symbol interference (ISI) as well as reflections and crosstalk may strongly degrade the signal
integrity along the transmission path. Finally also at the receiver side, a non-ideal equalizer (EQ)
and PLL inherent phase noise of the clock and data recovery (CDR) unit will additionally provide
their own timing jitter [3, 6, 8, 11, 15, 45, 53, 82, 83, 127].
A common way to highlight the problem of signal recovery and presence of jitter is the eye
diagram, as depicted in figure 2.2. It shows the received analog data signal folded in time, with
the bit period referred to as unit interval (UI). The untreated received eye is often almost closed
and has to be reopened with equalization techniques or signaling schemes that try to compensate
the channel influence. Timing jitter especially degrades system performance when causing a large
horizontal eye closure, since signal transitions spread over the entire bit interval and thus, impede
the recovery circuit to synchronize with the data. Inside the plotted waveform one may thus define
an eye mask [43, 47] which must not be violated in order to meet signal quality requirements.
5
2. F UNDAMENTALS OF J ITTER , PLL S AND BER A NALYSIS
Considering the timing budget or horizontal eye closure at the optimum decision threshold in
figure 2.2, simple probability distributions may be used for jitter analysis. Timing jitter is then
described as a statistical signal in terms of different components [82,104]. As shown in the scheme
in figure 2.3, an observed total jitter (TJ) distribution can basically be decomposed into a bounded
deterministic (DJ) and an unbounded random (RJ) part. Both components relate to independent
time-domain random processes and thus, appear as convolved in histogram domain.
Random jitter is usually considered as Gaussian, but can basically follow any unbounded prob-
ability behavior. It is observed at both distribution tails, extending them toward infinity. Bounded
deterministic jitter can be of arbitrary shape, and is expressed in terms of various subcomponents
in order to investigate and distinguish various root causes. DJ is further subdivided into sinusoidal
or periodic (SJ), bounded uncorrelated (BUJ), and data-dependent jitter (DDJ). Use of generated
SJ is especially important for jitter tolerance testing [43] and for the measurement of jitter transfer
functions [46, 121] in PLLs. Sometimes SJ also appears as an effect of parasitic spurs or power
supply noise. BUJ is mainly caused by couplings, such as crosstalk from adjacent transmission
lines, digital switching logic or ground bounce effects. BUJ is always considered bounded because
of the limited coupling strength. Exact models are difficult to derive for this component since both
coupling signals and mechanisms are highly variable. Finally, DDJ is a jitter component which
can be related to the transmitted data pattern. Duty-cycle distortion (DCD) is caused by a differ-
ence in the pulse width between logical high and low levels and ranges from voltage offsets or
different rise and fall times at signal transitions. Inter symbol interference (ISI), appears when
the channel impulse response extends over several bit periods. As shown in figure 2.4, a single
transmitted pulse is spread in time along the channel, and thus overlaps and influences adjacent
symbols. This causes a significant error in timing recovery. Fortunately, DDJ influences can be
fully compensated with an equalizer (EQ) if the channel is characterized by its impulse response,
or with an adaptive decision feedback equalizer (DFE) if it is unknown [5]. The use of EQs is re-
stricted on compensating DDJ, other jitter components which propagate through the transmission
path are then dealt by the receiver PLL.
6
2.1. J ITTER IN H IGH -S PEED S ERIAL L INKS
transmitted received
signal signal
0 1T 2T 3T 4T
In order to provide a fundamental understanding of the underlying research field, first the math-
ematical description of timing jitter and phase noise will be reviewed. In this context, jitter is seen
as time domain representation of phase noise, which describes the spectral purity of an oscillator.
Further, as a random process, phase noise has to be described in terms of statistical measures such
as variance and power spectral density (PSD). In the following sections thus, mathematical rela-
tions for these measures will be derived. These definitions and derivations are very common in
PLL literature and can for example be found in [30, 36, 42, 121].
This phase modulated signal can be decomposed in terms of Bessel functions. If the phase
variation of the noise signal is small compared to the reference period of the ideal oscillator
|φ(t)|1rad, equation (2.1) can be rewritten as:
The output spectrum of this signal consists of two Dirac impulses located at the carrier frequency
ω=±ω0 together with the frequency translated spectrum of φ(t). If φ(t) is considered as a station-
ary random process, its phase noise spectrum Sφ (∆ω) can be calculated by the Fourier transform
of the auto-correlation function. Here the variable ∆ω is used to denote the frequency offset from
the carrier ω0 , as shown in figure 2.5, and Sφ (∆ω) is referred to as double-sideband PSD, which
contains the spectral power of both sidebands of the oscillator spectrum. The phase noise is of-
ten quantified in terms of a single-sided spectral noise density L{∆ω}. This is the noise density
Sv (ω) Sv (ω)
∆ω 1−Hz
bandwidth
ω0 ω ω0 ω
7
2. F UNDAMENTALS OF J ITTER , PLL S AND BER A NALYSIS
measured at a frequency offset ∆ω from the carrier and is therefore one-half of Sφ (∆ω). L{∆ω}
additionally refers to the carrier power and is given in units of dBc/Hz:
Sφ (∆ω) noise power in 1 Hz BW at ω0 + ∆ω
L{∆ω} = = 10 log [dBc/Hz] (2.3)
2 carrier power
For a detailed description of the phase noise PSD also refer to [42, 78] or [30, chapter 11].
A typical phase noise spectrum for a voltage controlled oscillator (VCO) in high-speed PLLs
is given in figure 2.6. This spectrum is also known as Leeson process and consists of three dis-
tinct noise regions. The 1/ω 3 and 1/ω 2 terms represent flicker and thermal noise of electronic
components inside the VCO. These noise regions are integrated due to the frequency-to-phase
conversion, which corresponds to a multiplication of 1/ω 2 in spectral domain. The 1/ω 0 phase
noise floor is caused by external components and is not affected by the integration process.
The VCO noise model is usually specified using four parameters: the flicker noise corner fre-
quency ff l , a measured phase amplitude A1 with corresponding frequency f1 located in the 1/ω 2
region, and the noise floor amplitude AP hN . This phase noise model can also be realized in time
domain [122].
SΦ (∆ω)
∼ 1/ω 3
30dB/Dec
∼ 1/ω 2
20dB/Dec ∼ 1/ω 0
A1
0dB/Dec
AP hN
∆ω
ff l f1
Absolute Jitter jabs,k is the time difference of the k-th clock edge measured between an ideal
(tid,k ) and a non-ideal (tk ) clock signal:
Period Jitter jper,k is defined as the time variation of the clock period. It is the time difference
between k-th clock period Tk and the ideal period T :
jper,k = Tk − T
= (tid,k − jabs,k ) − (tid,k−1 − jabs,k−1 ) − T
(2.5)
= jabs,k−1 − jabs,k + tid,k − tid,k−1 − T
⇒ jper,k = jabs,k−1 − jabs,k
8
2.1. J ITTER IN H IGH -S PEED S ERIAL L INKS
(m)
Accumulated Jitter jacc,k is defined similar to period jitter, besides that the time displacement
of a non-ideal clock is measured m periods after the reference clock edge. According to this
(1)
definition we have jper,k =jacc,k .
(m) (m)
jacc,k = Tk − T (m)
(m)
(2.6)
⇒ jacc,k = jabs,k−m − jabs,k
T (2) T
1jacc,3
0 (2) 1
0
0
1 0
1
(2)
T3 T4 jper,5
0
1 0
1
jittery
clock
11
00 11
00 11
00 11
00 1
0
00
11 00
11 00
11 00
11 0
1
00
11 00
11 00
11 00
11 0
1
00
11 00
11 00
11 00
11 0
1
00jabs,1
11 jabs,2
00
11 00jabs,3
11 00jabs,4
11 0jabs,5
1
ideal
clock
Input-Output Jitter The most common method for analyzing the performance of a PLL is to
measure the time difference between input reference frequency and the PLL output clock. This al-
lows for directly quantifying the time domain misalignment of the PLL, which yields a qualitative
description for synchronization performance.
If the PLL is used for clock and data recovery (CDR) as with serial high-speed receivers, the
reference frequency is replaced by the analog input data and jitter values are measured between
bit transitions of the input signal and the PLL output clock (see figure 2.8). In order to correctly
quantify IO jitter, thus, an exact time interval measurement has to be performed.
IO jitter measurements are very important for practical use in high-speed communications and
required by a broad variety of applications, such as production tests, clock synchronization, or
signal quality specification [82].
9
2. F UNDAMENTALS OF J ITTER , PLL S AND BER A NALYSIS
{jabs,0 , T0 + jabs,1 , 2T0 + jabs,2 , . . . , kT0 + jabs,k }. Therefore, the k-th phase deviation caused by
jitter appears at time instant tk =kT0 + jabs,k , so that
In addition, if the absolute jitter is significantly smaller than one sampling period (jabs T0 ), we
have φ(kT0 + jabs,k )≈φ(kT0 ), and can rewrite equation (2.7) as
φ(kT0 )
jabs,k ≈ , (2.8)
ω0
where jabs,k is now a discrete time random process, which simply corresponds to a sampled and
scaled version of the continuous phase noise process φ(t).
The behavior can be summarized as follows: The VCO generates an output clock with frequency
fvco , which depends on the given input voltage. The phase detector compares the phase of this
clock against a reference frequency fref , and decides whether fref is preceded (early) or pursued
(late) by fvco . According to this decision, logical down (early) or up (late) pulses are generated.
Both signals drive a charge-pump which injects or unloads current into the loop filter, and thus
provides the control voltage for the oscillator. If the oscillator clock is late, several up pulses are
generated by the PD which increases the loop filter voltage and thus, moves the oscillator toward
higher frequency where both phases are again aligned. For an early oscillator clock the reverse
behavior is observed. This way a non-linear control loop acts as synchronization system.
In the past years, charge-pump PLL architectures have dominated the field of high-speed trans-
ceivers due to a low phase noise. Although all-digital PLLs [25, 110] are becoming increasingly
important with technology scaling and for design cost reduction, CPLLs are still widely used.
They offer two major advantages compared to pure analog architectures. First, a flexible design
can be achieved with decoupled parameters such as the loop bandwidth, damping factor and lock
range. Second, the included charge-pump allows for a zero static phase offset [46, 133].
In serial high-speed links both transmitter and receiver are characterized by PLLs as clock
synthesizers. The transmitter often uses an additional clock divider in the PLL feedback path to
multiply the reference frequency in order to yield the desired high-speed data rate. Conversely,
10
2.3. J ITTER AND BER A NALYSIS M ETHODS
the receiver is characterized by a clock and data recovery (CDR) circuit, where the serial input
data is used as reference frequency and the VCO output is the synchronized clock. The PD can
determine a phase mismatch only at input signal changes, which requires a sufficient amount of bit
transitions inside the data stream. Therefore, typically the 8b10b encoding scheme [43] is used,
which converts 8 bit of original data into 10 bit for transmission. This encoding guarantees for
sufficient bit transitions with a maximum of four consecutive equal bits, and a DC-balanced signal.
CPLLs have been analyzed and described thoroughly in literature, where the theory has been
extended from the linear model of analog PLLs [36]. A valid continuous-time approximation is
obtained, if the loop bandwidth is considered significantly smaller (at least 1/10) than the update
frequency of the phase detector. Since high-frequency signals are suppressed by the loop filter,
digital pulses of the phase detector are averaged and thus, a linear s-domain model can be used
for a CPLL design. At higher frequencies where the PD update rate is comparable to the loop
bandwidth, the feedback delay will induce an excessive phase shift and hence, lead to instability.
In order to account for this effect, discrete-time z-domain equations have been derived as well [46].
Unfortunately, these analytical equations still do not provide an accurate description of the non-
linear phase noise behavior inside a CPLL. Only behavioral time domain models that are able
to cope with the non-linear loop dynamics of a CPLL thus correctly reflect the true phase noise
of high-speed transceivers. Such a model will be implemented in chapter 8 to demonstrate the
application of a proposed jitter analysis method, and to guarantee that specification requirements
such as the target BER and jitter tolerance are met.
where B=10−12 is the desired BER level, a specifies the statistical probability or confidence level
that the true BER value is less than B, and E is the number of detected errors during measurement.
When no bit errors occur (E=0), the second term of the equation is zero and the solution to equa-
tion (2.9) is simplified. For example, with a=0.95 it is necessary to transmit N =3.0/B=3·1012
bits without errors in order to meet the imposed specification requirement. In a 3Gb/s transcei-
ver this would require an analysis time of T =N/3·109 =1000s=16m, 40s. Such a huge test time
cannot be tolerated for high volume production tests, where all specification requirements of the
transceiver have to be verified within several hundred milliseconds.
Therefore, test engineers have to rely on analysis methods which allow for accurate BER esti-
mation using a number of jitter samples which is several orders of magnitude smaller. Thus, cor-
responding mathematical models and equations must be provided in order to correctly verify the
11
2. F UNDAMENTALS OF J ITTER , PLL S AND BER A NALYSIS
desired target BER. Jitter values can usually be obtained easily from a model simulation, however
this process is often more complex in practice when carried out on hardware. High precision equip-
ment is required to perform accurate off-chip jitter measurements, including the use of high-speed
sampling scopes, time interval analyzers (TIAs) or bit error rate testers (BERTs) [12–14, 86, 130].
A detailed documentation of methodologies for jitter and signal quality measurements can also be
found in [43].
External noise sources can easily affect off-chip measurements at multiple Gb/s rates. Thus,
a broad variety of built-in jitter measurement (BIJM) systems [16–18, 35, 48, 57, 60, 65, 66, 100,
129, 131] has been developed in recent years as well. Such systems require a large amount of die
area if the jitter histograms have to be collected in real-time [16]. This is especially the case if for
example frequency domain analyses have to be realized and thus, all jitter values are needed. In
cases where the measurement time is uncritical, BIJM circuits also become very small but then,
they can only be used for histogram based jitter analysis.
Nevertheless, histogram based methods represent the most important class of analysis princi-
ples, since they directly relate jitter with the BER. This is not directly the case for time or frequency
domain based methods. In the following sections these three analysis domains will be explained
in more detail in order to give a comprehensive overview to the state-of-the-art in the topic.
12
2.3. J ITTER AND BER A NALYSIS M ETHODS
where T is the bit period, and tL and tR the resulting distances on the bathtub curve at 10−12 . The
total jitter peak-to-peak value TJpp or timing budget can thus directly be determined with
If normalized by the unit interval (UI) so that T ≡1, tL and tR equal the portion of eye closure.
These are the parts of the UI not accessible for sampling if the target BER has to be fulfilled.
The bathtub curve representation offers a simple way to verify whether a measured jitter distri-
bution achieves the specification. Unfortunately, in order to determine the correct eye aperture at
very low probability levels a huge amount of jitter samples must be collected. For BER levels of
10−12 and lower, a direct measurement of the histogram is not feasible. Especially in simulations
bathtub curves are only tracked down to probability levels that are orders of magnitude higher than
the target BER.
Therefore, an extrapolation of the bathtub curves is required. This extrapolation can be huge,
in the case of N =108 jitter samples it still ranges over four orders of magnitude, and can thus
only be done correctly if valid model assumptions are made for the underlying jitter distributions.
Common model assumptions are aligned to the popular Gaussian tail model [123] and can be
characterized as follows:
1. Jitter is a stationary random process.
2. The measured total jitter (TJ) distribution can be separated into two components, random
(RJ) and deterministic jitter (DJ).
DJ component
σL AL RJ component
Total Jitter AR σR
PDF
µL µR
−T /2 0 T /2
F IGURE 2.11.: RJ and DJ components of a jitter PDF. Definitions of right(R) and left(L) tails
correspond to the bathtub curves from figure 2.10.
According to these assumptions a TJ distribution can always be decomposed into two Gaussian
tails together with an arbitrary shaped bounded DJ component, as shown in figure 2.11. In order
to correctly extrapolate a measured distribution, analysis methods have to identify the three model
parameters µ, σ and A for both tails. This means, one is basically trying to fit a Gaussian function
into the measured distribution tails. Jitter analysis methods are thus also referred to as tail fitting
algorithms or jitter decomposition methods, while in mathematical statistics this problem domain
is also known as tail extrapolation and treated by extreme value theory. Once the model parameters
have been identified, the TJ timing budget can easily be calculated for arbitrary probabilities and
thus used for BER analysis. A mathematical description of the timing budget is provided later on
in section 3.
Various methods were developed to separate the random and deterministic jitter components
with tail fitting algorithms [51, 54, 58, 84, 95, 124, 136]. In this section, existing techniques are
reviewed in order to provide a comprehensive overview to the state-of-the-art.
13
2. F UNDAMENTALS OF J ITTER , PLL S AND BER A NALYSIS
14
2.3. J ITTER AND BER A NALYSIS M ETHODS
of distribution tails, and can also be used to derive a unifying optimization scheme for tail fit-
ting which covers all of the above described references [51, 54, 95, 111, 123]. Further, it can be
generalized for use with arbitrary non-Gaussian tails.
Other Methods
Other less popular jitter decomposition methods are based on techniques for deconvolution [124,
125,128], Gaussian mixture models [98] and the wavelet transform [136]. Deconvolution methods
rely on the idea that in histogram based analysis a total jitter PDF is given as convolution result of
the RJ and DJ components. If one of these two components is approximately known or estimated,
a deconvolution algorithm can be used to determine the other component, and thus to retrieve the
Gaussian model parameters. A major drawback of these methods is that they suffer from accuracy,
since either the DJ or RJ component has to be estimated prior to the deconvolution.
Another method is based on the wavelet transform [136] and uses derivatives of Gaussian wave-
lets to detect the locations (mean values) of the Gaussian model functions. The variances are deter-
mined from a transformed log-likelihood function, while Gaussian amplitudes are not considered.
Due to the applied wavelet transform, this approach also suffers from a high computational de-
mand.
15
2. F UNDAMENTALS OF J ITTER , PLL S AND BER A NALYSIS
In [55, 56] four spectral regions of the jitter transfer function of CDR circuits are defined to
allow for BER analysis. The approach is restricted to Gaussian RJ combined with SJ, where the
sinusoidal jitter frequency is extracted from the spectral information. The obtained jitter transfer
characteristic of CDRs is subsequently [101, 102] also used to derive an analytical approximation
for the maximum phase error, which can be adapted for BER calculations.
In his book [82] Li thoroughly describes frequency domain principles for jitter separation in-
cluding DDJ, SJ and RJ types. BUJ can generally not be separated from the RJ noise floor, unless
it can be measured independently or controlled in some way.
16
3. Mathematical Background
This chapter deals with the mathematical basics involved with the developed jitter analysis meth-
ods and optimization schemes. First, the quantile normalization is reviewed as fundamental math-
ematical principle for a powerful class of tail fitting methods which is going to be analyzed and
optimized throughout subsequent chapters. Then performance metrics and test distributions are
introduced for the qualitative analysis of tail fitting methods.
This function allows to represent a distribution by the quantile plot (or QQ-plot) [19, 20, 116]:
−1 i
F , x(i) : i = 1, . . . , N (3.3)
N +1
where observed amplitudes x(i) are represented in terms of amplitudes of the theoretical model
quantiles F −1 (pi ). For a large sample size N , the sample quantiles x(i) approximate a shifted
and scaled version of the theoretical ones [115, 116]. This offers a linearized perspective on dis-
tributions, which is especially useful for tail fitting. As an example, in figure 3.1 an empirical
distribution F̃ (x) is shown, obtained from N =100 random samples of a normal distribution with
zero mean and unit variance F (x)=N (0, 1) (solid line). The quantile function F −1 (p) at the right
transforms the empirical distribution into an easy-to-fit linear function.
Thus, if a gathered distribution exactly matches the expected probability function as described
by the inverse F −1 , the result is a perfect line along the unit diagonal. If for example only the
tail part follows an expected behavior, as is the case for the RJ-DJ model with Gaussian tails
17
3. M ATHEMATICAL B ACKGROUND
1
2
0.8
1
0.6
Q(pi)
0
pi
0.4
−1
0.2
−2
0
−2 −1 0 1 2 −2 −1 0 1 2
xi xi
(a) (b)
F IGURE 3.1.: Example for a) an empirical distribution function CDF(x)=F̃ (x) with N =100
random samples of N (0, 1) and b) the corresponding quantile plot.
(section 2.3.1), this line is still observed at the tail parts. This way, the tail fitting problem can
basically be simplified to a linear regression analysis.
Such a linearizing transform offers a great simplification for the tail fitting procedure, especially
in terms of computational demand. As we will see in chapter 4, the method of least squares can be
implemented very efficiently for this purpose, as it uses only summing terms and recursions. Fur-
ther, the residual error after transform becomes approximately constant and normally distributed
over a large probability region, which makes the least squares method an ideal candidate for the
maximum likelihood estimation of tail parameters.
where µR is the mean value and σR the standard deviation of the Gaussian. The parameter ρT is
the transition density and reflects the probability of bit transitions in the transmitted data signal.
In a clock-like ‘1010. . .’ pattern for example we have ρT =1, while for pseudo random binary
sequences (PRBS) ρT =0.5. In our case the BER definition describes the probability course p of
the right bathtub curve with negative jitter values and thus, represents the right sided cumulative
18
3.1. Q UANTILE N ORMALIZATION
BERR (x)
F (x) = p = CDFR (x) = (3.5)
ρT
In order to obtain a standardized representation of the integral in equation (3.4) the variable q
normalizes a Gaussian function with respect to mean µ and standard deviation σ:
x−µ
q= (3.6)
σ
With the standardized variable we yield
q q 02
1
Z
CDFR (q) = p = √ e− 2 dq 0 (3.7)
2π −∞
For the left bathtub curve CDFL (x)=1−CDFR (x) the same result is obtained, when using the
negative BER integral from x to ∞. The inverse Gaussian probability function or quantile norma-
lization is thus given by
−1
√
Qgauss (p) = q = Fgauss (p) = − 2 · erfc−1 (2 · p) (3.10)
In jitter analysis Qgauss (p) is briefly known as the Q-function [54, 82, 123]. It is commonly used
to transform measured probability functions into Q-domain, where tails appear as straight lines.
Often Qgauss (p) is defined using a positive sign. Here, the negative sign is used explicitly to
maintain symmetry between probability domain and Q-domain.
An example for a Q-normalized bathtub is given in figure 3.2. The linearizing effect on Gaussian
tails yields curves which can easily be fitted and extrapolated by simple linear functions. The
standardized variable q as defined in equation (3.6), makes the quantile normalization independent
from mean and standard deviation of the original Gaussian model. As a direct consequence, the
mean value µ is mapped onto a line offset, while the standard deviation σ is mapped onto a line
slope in Q-domain. After the transform, both parameters can easily be retrieved from the linear
regression, as also shown in figure 3.2. The zero crossings of the lines correspond to the Gaussian
means µL and µR , while the standard deviations σL and σR are given by the respective tail slope.
A coefficient comparison between standardization and obtained linear function yields:
q = (x − µ)/σ ⇐⇒ q = o + s · x
19
3. M ATHEMATICAL B ACKGROUND
CDF(x)
10−6
10−9 teye
10−12
tL (10−12 ) tR (10−12 )
x
slopes
Q CDF(x)
−1 µL 1/σL 1/σR µR
−3
−5
−7
x
0 T
F IGURE 3.3.: Simple optimization scheme for Gaussian tail fitting based on Q-normalization.
line offset o and slope s. These values are then used to retrieve the Gaussian tail parameters µ and
σ. This optimization has to be carried out for both distribution tails independently.
Note, that the presented optimization scheme does not consider the Gaussian tail amplitude
A as a third model parameter. This forms a missing gap for many tail fitting methods based on
conventional Q-normalization [51, 54, 82, 123]. Therefore, Miller [95] introduced an additional
variable for amplitude normalization. In chapter 4 a way to include this variable into the present
scheme is shown, which closes the missing gap and significantly improves fitting performance.
Returning to the recovered Gaussian model parameters µ and σ, the TJ timing budget as im-
portant quality measure for high-speed interfaces can now be easily retrieved. According to the
bathtub function in figure 3.2, the peak-to-peak value of total jitter TJpp is decomposed into RJ
and DJ:
TJpp = DJpp + RJpp (3.12)
where each of these components is described in terms of the tail parameters:
DJpp = µL − µR (3.13a)
σL + σR
RJrms = with RJpp = RJrms · −2 · Q(BERspec ) (3.13b)
2
Here, RJrms denotes the root-mean-square value and is calculated as the mean of the two Gaussian
standard deviations. The probability dependent Q-factor from equation (3.10) denotes the units of
Gaussian standard deviations we have to move away from the mean value, in order to reach the
desired target BER level. With BERspec =10−12 , equation (3.12) is rewritten in the commonly
used form:
TJpp = DJpp + 14.07 · RJrms (3.14)
In order to highlight the influence of RJrms on the total amount of jitter, in table 3.1 important
probability levels together with their Q-factors are given.
20
3.2. P ERFORMANCE A NALYSIS OF A LGORITHMS
BERspec −2 · QBER
10−6 9.51
10−9 12.00
10−12 14.07
10−15 15.88
TABLE 3.1.: Multiplicative constant to specify a target BER for TJpp values.
A general problem appears with fitted regression lines in Q-domain. The Q-tails often approach
the linear behavior asymptotically, which can end up in misleading TJpp estimates if the fit is
not performed in a suitable probability region. Due to the asymptote it is not possible to detect
an exact probability level where the linear behavior begins. With the proposed fitting method in
chapter 4 this effect is also visible and has to be investigated carefully in order to afford accurate
tail extrapolations. In sections 4.3.2 and 4.4 this problem domain will especially be addressed by
focusing on performance optimizations with additional fitting parameters.
Here, K is the number of evaluation runs and should at least equal a few hundred, in order to
construct empirical relations. Sometimes a tail fitting method might also produce misleading
outliers due to convergence problems. In this case it is better to use the median value Emed
together with the interquartile range IQR (interval between upper qup and lower qlo quartile) to
specify statistical spread. These measures are less prone to outlier degradation [29]:
21
3. M ATHEMATICAL B ACKGROUND
According to this definition, EL defines a positive error threshold, which is exceeded by less than
(1−a)/2 ≈ 2.2% of estimates.
Sometimes a tail fitting algorithm may also produce misleading outliers, especially when con-
vergence failures occur. Such failures yield error distributions that are quite different from the ideal
Gaussian as shown in figure 3.4, and are typically characterized by slowly decaying heavy tails.
Outliers have to be avoided as far as possible. In order to measure their presence, the fourth stan-
dardized moment or kurtosis κ can be used. This statistical moment describes the “peakedness”
of a distribution and yields a value of κ=3 for an ideal Gaussian. If a distribution is outlier-prone,
κ will be significantly larger, while for the bounded uniform case it is κ=1.8.
22
3.2. P ERFORMANCE A NALYSIS OF A LGORITHMS
Quad. Curve DJ TJ
Triangular RJ
σRJ
Sinusoidal
Uniform
ADJ
F IGURE 3.5.: Test distribution types, constructed with parameters σRJ and ADJ
zero mean and standard deviation equal σRJ as developed in [10], while DJ PDFs are constructed
with bounded random processes according to the respective DJ shape. Sinusoidal jitter additionally
uses a sinus function of random phase ϕ and frequency fSJ . The uniform process generates a
random number out of the bounded interval [−ADJ /2, +ADJ /2], while triangular and quadratic
curve shaped jitter can be realized as two or three superimposed uniform processes. The total jitter
random process JT J is simply obtained by the addition of RJ and DJ components:
where the random processes correspond to the shape characterizations from table 3.2.
Note that the distribution synthesis with the two components JRJ and JDJ in time domain yields
a convolution in histogram domain. Due to the addition of independent random variables, the
resulting TJ distribution will be decomposed into DJpp and RJrms , which differ from the original
σRJ and ADJ parameters. According to the central limit theorem, the combination of an arbitrary
bounded random process with an unbounded Gaussian process always leads to a distribution which
is more Gaussian-like than the prior bounded component. Thus, an increase of the RJ component
(RJrms ≥ σRJ ) as well as a decrease of the DJ component (DJpp ≤ ADJ ) will be observed [123].
For sinusoidal DJ type, this topic has also been investigated in the appendix of [52].
The resulting TJ shapes can be characterized in a representative way using the variable ratio
σRJ /ADJ . The TJ shape then depends only on the relative difference between the two variables,
while the distribution size can be described by just one of them.
The TJpp,true values at the target BER are determined using numerical approximations. There-
fore, closed form equations of the independent RJ and DJ components are convolved, which allows
for a direct approximation of the complete TJ shape. Then, the TJ value closest to the 10−12 level
23
3. M ATHEMATICAL B ACKGROUND
is extracted and an additional Newton step carried out. The relative numerical error is guaranteed
to be smaller than 10−4 .
Throughout subsequent analyses, the specified test distributions will always relate to the inde-
pendent parameters σRJ and ADJ prior to convolution. This is to provide reproducible simulation
results. The uniform DJ shape will especially be utilized as a reference, since it represents a good
compromise between easily decomposable sinusoidal shape, and hardly separable triangular or
quadratic curve shapes.
24
4. A Fast and Accurate Jitter Analysis
Method
With the mathematical background from the previous chapter a novel, fast and accurate jitter
analysis method is developed. This method puts the Q-normalization into the context of a com-
plete three-dimensional Gaussian model optimization, where the unknown tail parameters mean
µ, standard deviation σ and amplitude A are identified for both distribution tails. The optimiza-
tion scheme is realized with simple recursions that allow for a very fast exploration of the search
space. As will be demonstrated in this chapter, the proposed method yields very accurate fitting
results combined with low computational demand and a flexible design architecture. It automat-
ically determines the best suited tail part for distribution tail fitting, and thus represents a clear
improvement to existing tail fitting methods.
The three-dimensional approach is based on an additional scaling factor, included with the op-
timization scheme from figure 3.3. Thus, it also allows for tail amplitude search. Although the
proposed principle has been developed independently, this idea is not novel. Popovici [111] al-
ready described the mathematical principle for Gaussian quantile normalization with respect to
unknown amplitude A and standard deviation σ. There was no need to include the mean value
µ, since the signal to noise ratio was used for BER analysis and the application focus was not on
jitter decomposition. Miller [95] was the first to suggest the amplitude scaling factor which finally
allowed for three-dimensional Gaussian tail fitting. Unfortunately, both the determination of the
tail part as well as the optimization scheme for model parameter search were not described.
In this chapter a complete approach to three-dimensional tail fitting based on Gaussian quantile
normalization is provided. A simple optimization scheme is first derived, where the search algo-
rithm simultaneously searches for the Gaussian tail part while identifying the best suited model
parameters. A detailed and thorough description of the algorithmic characteristics and associated
mathematical fundamentals outlines an excellent fitting quality. A comprehensive performance
analysis then gives an impression on the potential of the proposed method. It involves the com-
bination of different test distributions and sample sizes. Subsequent performance optimizations
further improve estimation accuracy as well as the robustness of the algorithm. Therefore two
conservative concepts based on probability domain parameters as well as a Q-domain threshold
are proposed. After respective performance analyses with optimized parameters, the chapter con-
cludes with a summary of the novel jitter analysis method.
Due to the excellent tail fitting performance combined with a fast implementation, flexible ar-
chitecture and a minimum of conservative fitting parameters, the developed method is also meant
to act as a reference for future designs. For this purpose, in chapter 6 a comprehensive perfor-
mance comparison is carried out against other methods based on quantile normalization. Since the
tail fitting method utilizes the quantile normalization in a scaled sense, throughout this thesis it is
referred to as “scaled Q-normalization” (sQN) method.
25
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
σ, and amplitude A, best matching the measured distribution tails. As was already shown in
section 3.1.2, the quantile normalization of Gaussian tails yields linearized curves, which can be
used for regression analysis. Along these tails in Q-domain the two parameters µ and σ can easily
be retrieved, but not the amplitude A. Now the same principle is extended in order to determine
all three Gaussian model parameters.
Considering a typical jitter distribution with the PDF as shown in figure 4.1 (solid curve), the
probability function or CDF covers the complete probability range from zero to one. Thus, the
PDF area equals A=1. Similar, a pure Gaussian function N (x) (dashed curve at the left) with
same area A=1 can be fitted into the distribution tail. For the moment it is assumed that the left
Gaussian function N (x) represents the ideal fitting result with known tail parameters µ and σ.
PDF(x)A=1 PDF(x)A=1
→
N (x)A=1 N (x)A<1
At the left of figure 4.1 a comparably small part of the Gaussian function overlaps the outer
left tail of the jitter PDF. This means, the Gaussian function with A=1 cannot be fitted nicely into
the distribution tail, even with known parameters µ and σ. A smaller Gaussian amplitude instead,
would allow for an optimized fit with respect to both tail length and fitting error, as depicted in the
right figure. This means, an adapted Q-normalization function must be found to optimize the fit,
so that the Gaussian model best matches the distribution tail. With different jitter PDFs, it would
theoretically be necessary to derive an adapted Q-normalization function for each of the possible
tail areas A < 1, which is not feasible.
Instead one can think of a reverse approach where the probability of distribution samples is
scaled by a multiplication factor k, and the Q-normalization remains constant. This principle is
demonstrated in figure 4.2, where the original PDF is blown up by the scaling factor k. Although
the obtained probabilities will obviously increase and thus, span an area which is larger than one,
the tail fitting principle now has to be seen from the perspective of the constant Q-normalization
stage. This means, the normalization is narrowed down toward a smaller probability region, be-
cause of the scaling. In fact, the k-scaled distribution is Q-normalized in an original probability
region from zero to 1/k. Thus, the scaled distribution corresponds to a Gaussian tail search with a
smaller area of A=1/k.
PDF(x)A>1
PDF(x)A=1
→
N (x)A=1 N (x)A=1
26
4.1. S CALED Q-N ORMALIZATION
From the third expression we can see, that RJpp now also depends on the amplitudes of the Gaus-
sian models. For the special case AL =AR =1, the Q-function yields Q(BERspec ) and the model
reduces to the same equation as in (3.13).
In the following subsections the proposed scheme is described in more detail. First, the focus
is put on an efficient realization of the optimization procedure and the involved search algorithm.
Then, the generalization property of the scheme is highlighted as an additional feature. An error
analysis is carried out to justify the quantile normalization with associated linear regression as
a fundamental mathematical principle, which is close to the optimum solution for the tail fitting
problem. Finally, also the algorithmic details for an implementation in C++ are provided.
27
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
0
10
2
k=16
k=4
k=64
0 −4
10
−2
−8
10
−4
−12
10
0.78 0.81 0.84 0.87 0.7 0.8 0.9 1
Sampling Time [UI] Sampling Time[UI]
(a) k-scaled Q-functions for k={1, 4, 16, 64}, the maxi- (b) Left bathtub distribution with fitted Gaussian tail.
mum linearity is obtained at k=7.15
F IGURE 4.4.: Scaled Q-normalization principle demonstrated with N =107 jitter samples of
a 1 Gb/s signal (measured with Agilent Infiniium 40 GS/s). The jittery data
is generated with a SyntheSys BERT-Scope 7500A (sinusoidal ADJ =0.3 UI,
σRJ =0.2/14.07 UI), which also provides the measured magenta bathtub samples
at lowest BER level in (b). The required calibration delay is chosen to match the
highest BERT point with the measured Agilent bathtub.
This function may be plotted for various values of k as shown in the example of figure 4.4, where
a typical jitter distribution is analyzed. The effect on the original CDF is observed as bent Q-
functions, which achieve best linearity for a certain scaling factor. Here, the resulting analysis
domain is also referred to as scaled Q-domain, where the optimum scaling factor yields a linear
function which is best described by a Gaussian. The obtained fitting result already shows the
potential of this approach, when performing tail extrapolations over several orders of magnitude.
After Q-normalization, the second optimization stage fits a regression line into the tail region
of the k-scaled Q-function, using the method of least squares. It yields fitted slope s, offset o
and regression error σ̂err , with the error as fitness measure for the optimization procedure. An
essential speed-up is achieved by choosing a representation with the fitting length n as variable.
This variable denotes the n outermost points on a bathtub tail which are used for regression
analysis. A finite time resolution for jitter values is used, so that distributions become discretized
and consist of a limited amount of R bins per UI. One can also think of dividing the bit period
into equally sized steps. This leads to an adjustable time resolution 1/R, which greatly reduces
computational demand. Instead of each single jitter value, the linear regression is now carried
out only along the reduced number of bins. When chosen too small, R will obviously degrade
the fitting performance. In simulators this can be avoided by selecting R sufficiently large, but
hardware systems will usually suffer from coarse resolutions. This problem domain is especially
addressed in chapter 5, when focusing on hardware design aspects.
Each collected jitter value is assigned to a bin of the discretized distribution function. The
regression error σ̂err can thus be represented as a function of the number of fitted bins n, and sub-
sequently used for optimization. In figure 4.5, σ̂err is plotted as a two dimensional function of scal-
ing factor k and tail length n, demonstrating its usability as goodness-of-fit measure. The global
28
4.1. S CALED Q-N ORMALIZATION
0
10
Regression Error
−1
10
−2
10
0
10
1
1 10
10
2
2
10
Scaling factor k 10 Regression length n
F IGURE 4.5.: Regression error σ̂err depending on fitting length n and scaling factor k. The
example is the same as in figure 4.4 with a time resolution of 1.83 ps (1 UI =
1000 ps).
error minimum is obtained at optimized k and n, and can be used to retrieve the three Gaussian
model parameters. Although the error is given as a function of only two variables σ̂err =f (k, n),
Gaussian tail fitting remains a three dimensional optimization problem in a strict mathematical
sense. The variable n only hides the linear regression, which deals with the two parameters line
slope s and offset o.
The linear regression stage detects the error minimum by recursively incrementing n over the
Q-tails. That is, the search algorithm starts with a few outermost tail samples or data pairs (xi , qi )
in Q-domain, and moves toward higher probability levels by recursively adding samples from the
bins. Thus, for each additional sample the investigated Q-region becomes larger, while qualita-
tively described by the corresponding regression error σ̂err (n). This procedure has two major
advantages. First, all the outermost samples are included, which is very important for tail fitting as
they belong to the Gaussian tail part. Second, the linear regression offers very simple recursions
for adding tail samples, and hence, the desired error minimum is detected very efficiently.
With given data pairs (xi , qi ), regression analysis assumes the linear relation:
qi = o + s · x i , i = {1, . . . , n} (4.4)
The regression coefficients and the error are calculated using least squares equations [20, p. 393]:
P P P
n · x i qi − x i · qi
s= P 2 (4.5a)
n · x2i −
P
xi
P P
qi − x i · s
o= (4.5b)
rP n
(qi − o − s · xi )2
σ̂err = (4.5c)
n−2
where s and o are the estimated parameters for line slope and offset, and σ̂err is the standard error
to be minimized. Since n is constantly incremented during optimization, the present summing
terms can be implemented very efficiently as recursions.
29
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
In order to determine a global error minimum, the linear regression stage must be applied to
every value of k. To avoid high computational load, a logarithmically scaled search grid is utilized
for initial estimation of k. In a refined minimization an accurate estimate is then obtained after
a few more iterations. Once the optimization process is completed, the Gaussian tail model is
simply given by the fitted parameters:
The factor k forms the reciprocal of the Gaussian amplitude as already described previously. The
parameter σ is the reciprocal of the gradient or slope of the linearized Q-tail. This is due to the
inherent property of the Q-function, to normalize a given distribution in units of Gaussian standard
deviation (see equation (3.6)). Finally, µ is the jitter magnitude where the regression line crosses
the zero value in Q-domain and decomposes the jitter distribution into bounded DJ and unbounded
RJ components. Thus, with
!
q = 0 = µ · s + o ⇒ µ = −o/s (4.7)
we yield the third expression in equation (4.6) and hence, confirm the result obtained with the
comparison of coefficients in equation (3.11).
The proposed scaled Q-normalization method must be applied to both distribution tails sep-
arately. For negative jitter values the right sided distribution function CDFR (x) is used as in-
put to the optimization scheme in figure 4.3, while for positive jitter values the reverse function
CDFL = 1 − CDFR is utilized. The obtained left and right Gaussian tail parameters are finally
able to decompose a measured jitter distribution into RJ and DJ, as described by equation (4.2).
As will be demonstrated in subsequent performance analyses (section 4.2), the presented ap-
proach achieves excellent accuracy, even if a comparable small amount of jitter samples forms the
distribution. Due to the linearization of a Gaussian function in Q-domain, the three-dimensional
optimization problem (µ, σ and A) is basically simplified to a linear least squares regression with
preceding data normalization. A key advantage is the efficient application of recursions inside
the regression stage, making the optimization process very fast. Other approaches, such as fit-
ting algorithms based on chi-squared tests [52, 84, 90] have to face a non-linear three-dimensional
optimization. Hence, they are complex and suffer from a high computational demand.
CDF(x) Error
Linear Reg.
×k F −1 (. . .)
s, o
A σ, µ
30
4.1. S CALED Q-N ORMALIZATION
which corresponds to the inverse CDF of unit amplitude, unit standard deviation and zero mean.
The corresponding scheme is given in figure 4.6.
Possible candidates for tail fitting other than the Gaussian model are distributions that exhibit
power-law behavior at the tails, such as the generalized Gaussian, generalized extreme value or
generalized Pareto distributions. These functions include large classes of tail shapes and intro-
duce additional degrees of freedom to the optimization scheme. In chapter 7 this generalization
principle will be further discussed.
The ability to substitute the Gaussian quantile normalization with any desired tail shape makes
the proposed scheme very flexible and thus, a powerful approach to tail extrapolation. In addti-
tion, if the amplitude pre-scaling factor is omitted, it becomes fully consistent with prior jitter
decomposition methods based on conventional Q-normalization as described in [51, 54, 82, 123].
ri = qi − q̂i = qi − o − s · xi , i = 1, . . . , N (4.9)
A residual plot depicts ri against fitted values q̂i and thus, visualizes trends or non-constant error
in the scatter behavior of residuals which is also denoted as heteroscedasticity [20, p. 340]. Ideally,
ri should be randomly scattered over the whole plot. This guarantees for a constant error variance
which is independent from the fitted model value and thus, fulfills the first of the conditions above.
However, for quantile normalization this is not the case in the outermost tail region, as shown in
the example of figure 4.7. Here, the residuals of 25 normal distributions with standard deviation
σRJ = 0.1 UI, zero mean as well as a sample size of N =104 are represented in a scatter plot. Due
to the known parameters, fitted values q̂ were replaced by the sample amplitude x, to highlight the
error structure along the unit bit period.
The reason for the observed heteroscedasticity is the limited sample size with probability gran-
ularity 1/N . This effect can be visualized by calculating the confidence bounds for quantiles. The
31
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
0.6
0.4
a=0.99
0.2
Residuals
0
−0.2
a=0.01
−0.4
−0.6
−0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4
x
F IGURE 4.7.: Scatter plot of residuals for 25 realizations of a normal distribution with sample
size N = 104 and σRJ = 0.1 UI. The curves correspond to the 1%, 5%, 10%
and 20% upper and lower confidence bounds when transformed into Q-domain,
according to equation (4.12).
probability of an observed quantile to be smaller than the theoretical one corresponds to a binomial
random variable with parameters (N, i) [71, sec. 3.1] [117]. Using the binomial distribution this
yields:
i
X N j
a = P (X(i) ≤ xp ) = p (1 − p)N −j = I1−p (N − i, i + 1) (4.10)
j
j=0
where X(i) is the i-th order statistics, xp the theoretical quantile at probability p, and a ∈ [0, 1] the
confidence level. Further,
Γ(u + v) x u−1
Z
Ix (u, v) = t (1 − t)v−1 dt (4.11)
Γ(u)Γ(v) 0
is the regularized incomplete beta function [31], which is used for numerical calculations. The
confidence level a, sample size N and probability p of the target quantile xp = Q(p) are fixed, and
we are searching for i = f (a, p, N ) such that equation (4.10) is fulfilled. The residuals ri,a for the
Gaussian distribution example are finally given by
where Q is the quantile function, and the resulting ri,a (p) are plotted in figure 4.7 for different a
values. At a=0.5 the residual function follows the zero line.
A possible way to compensate existing heteroscedasticity is to use a generalized least squares
approach, where observed quantiles are weighted according to their variance. However, even
with known variance this is difficult, since quantiles are highly correlated. Thus, the complete
covariance matrix Σ must be described. According to [71, sec. 4.8] the covariance elements σij
can be determined as:
min(pi , pj ) − pi · pj
σij = (4.13)
f (Q(pi )) · f (Q(pj ))
32
4.1. S CALED Q-N ORMALIZATION
where pi are the quantiles as defined in equation (3.1), f = F 0 denotes the probability density
function and f (Q(p)) is known as sparsity function [106]. An approximation of Σ is for example
given in [117] for the case of generalized extreme value distributions. The presented approach also
highlights the matrix computations involved with generalized least squares, and yields a regression
model with uncorrelated, constant errors. However, the inverse Σ−1 must be calculated, which
is only feasible for small sample sizes. This is an essential drawback which impedes utilizing
least squares equations (4.5) as efficient recursions. Nevertheless, heteroscedasticity of Gaussian
quantiles only influences the outermost tail region. With the huge sample sizes given in jitter
analysis scenarios, its influence on fitted tails becomes sufficiently small for a majority of test
cases, as will also be demonstrated later on.
Note, that other tail fitting methods that are not based on the quantile normalization as lineariz-
ing transform, have to face a non-linear regression where the error structure may highly degrade
the quality of fit. The Q-normalization technique instead, can at least always guarantee for an
approximately constant error in a higher probability region.
The third optimality condition as listed previously, demands x to be observed or measured with-
out error, which can also be guaranteed only for simulations but not for hardware measurements.
In section 4.1.1, already a time resolution variable was introduced, to speed-up the optimization
process using distributions with a discrete number of bins R. Thus, collected jitter values addi-
tionally suffer from a rounding effect or quantization into integer multiples of 1/R. This causes
an error in x, which also degrades the quality of quantile normalization. Its effect on fitting per-
formance will be investigated thoroughly in chapter 5.
In order to maintain efficiency of the scaled Q-normalization method, the analysis focus is
only on performance optimizations where the proposed scheme can utilize the fast least squares
recursions. Thus in sections 4.3 and 4.4, optimization will especially be carried out by introducing
conservative tail fitting parameters or by selecting suitable goodness-of-fit measures for the search
routine. In this context for example, an important question to be answered is, whether outermost
tail samples should be discarded or not.
33
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
Read parameters
from options file Qk =Q(CDF(x) · k)
*.opt
Linear Regression
generate σ̂err , o, s
random jitter value
store as σ̂err (n)
store into PDF
of resolution R
n++
reg. length
no Qk (n)≥0?
jitter sample
loop yes
N×
σ̂n =min{σ̂err (n)}
scaled k=k · ∆k
Q−normalization
scaling factor
method
no k≥kmax ?
store TJpp value yes
F IGURE 4.8.: Flow graphs of implemented testbench (left) and tail fitting algorithm (right).
Jitter samples are generated according to the involved time domain random processes as al-
ready described in section 3.2. These random processes are realized as C functions from [34], as
an improved alternative to the built-in standard C library. A Mersenne twister is used for gen-
erating uniformly distributed samples and the method from [10] transforms them into a normal
distribution.
Generated jitter values are assigned to a PDF vector which uses a discrete amount of bins R,
specified in the options file. The PDF is thus represented by an integer vector where each bin
represents a discrete time interval. When a jitter sample falls into a certain time interval, the
corresponding vector entry is incremented, equal to a counter variable. To obtain the simulated
probability values, counter values only have to be normalized by the sample size N . This principle
limits data memory when gathering large amounts of jitter samples, and allows for modeling
the limited time resolution of hardware systems. Later on, with PLL behavioral simulations the
resolution of the simulator will be selected as 1fs. For the investigated 3Gb/s serial interface
34
4.1. S CALED Q-N ORMALIZATION
this yields Rsim =3.33·105 bins per bit period, which guarantees for a sufficiently detailed timing
resolution so that quantization effects can be neglected.
Once N jitter samples have been gathered, the CDF is calculated as integral of the jitter distri-
bution. The scaled Q-normalization method then fits Gaussian functions into the distribution tails,
and returns the estimated timing budget TJpp,est at the desired target level of BERspec =10−12 . In
order to allow for a statistical analysis of this timing budget, K evaluations are carried out. The
described key analysis parameters are also summarized in table 4.1, together with their default
values.
After the evaluation loop, simulation results are stored in two separate output files. A logfile
(*.txt) contains recorded information about simulation progress, successful termination, start and
stop time, as well as a copy of the input options file. The output data file (*.dat) contains all
estimation results, such as the timing budgets over multiple evaluation runs. This file is meant to
be used by MATLAB scripts for post-processing and representation of results.
Fitting Algorithm
The implementation of the tail fitting algorithm is depicted in the right flow graph of figure 4.8. It
represents the realization of the scaled Q-normalization block at the left. As already described in
section 4.1.1 the basic algorithm is implemented as an optimization procedure which minimizes
the fitting error σ̂err as function of regression length n and scaling factor k. This concept is
realized with a nested loop for the regression length and an outer loop for the scaling factor.
The algorithm uses an initial search grid to identify the best suited scaling factor k, and hence,
starts with the first scaling value at k=1. With the CDF as input, the k-scaled Q-function is calcu-
lated and the linear regression analysis performed. The nested loop collects error values σ̂err (n)
over increasing regression length and continues until a maximum value of Qk ≥0 in Q-domain is
reached. This limit corresponds to a scaled CDF probability of 0.5 or half of the Gaussian tail
model. It aids in excluding CDF samples which hardly belong to the measured Gaussian tail, and
avoids negative influence of unfavorable DJ shapes.
The minimum error value of the nested loop is the optimum along the first search dimension n,
and is stored in a vector according to the different scaling factors of the outer loop. The second
search dimension k uses a logarithmically scaled search grid with the grid distance ∆k=1.2 as
default value. The parameter kmax =1/Amin must be chosen to include the minimum expected
tail amplitude, and thus, also affects computational demand of the search algorithm.
The coarse minimum along the second search dimension k is given by the initial search grid,
which must be further refined. This refinement process is carried out with a C implementation
of the MATLAB function fminbnd() from the Optimization Toolbox [88]. The required upper
and lower search bounds are the adjacent grid values of the selected scaling factor ∆k, with the
final result located inside the interval [k/∆k, k · ∆k].
After the refinement, the optimization of σ̂err is concluded, and the Gaussian model parameters
can be retrieved from the resulting kopt and nopt values together with the regression coefficients
35
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
in equation (4.6). Note that the tail fitting algorithm at the right hand side of figure 4.8 has to be
applied to both tails of a jitter distribution. The total jitter estimate TJpp,est is finally calculated
with both Gaussian model parameters according to equation (4.2).
In the C++ environment where the fitting algorithm has been implemented, the computational
effort depends on various factors. First, the Q-normalization function utilizes the inverse comple-
mentary error function erfc−1 (x), which can only be solved numerically and is computationally
expensive when using iterative approaches. As a solution, the MATLAB erfcinv() function
offers a very fast polynomial minimax approximation, with a relative error ≤1.13 · 10−9 . This
function has been transferred to the C++ environment. Second, the linear regression analysis has
to be carried out over the complete distribution tail inside the nested loop. This evidences the
importance of equations (4.5) which use a recursive solution for calculating line offset, line slope
and regression error.
Finally, the time discretization into R number of bins per UI highly influences computational
effort as it also defines the number of nested loop iterations for a given jitter distribution. In fig-
ure 4.9 the average calculation time tc is determined for a typical test distribution which occupies
approximately half of the unit interval. With a larger R, test distributions contain more bins and
thus, the computational demand is linearly increased. Simulations are carried out with an Intel
Core Duo 2.2GHz laptop, where the scaled Q-normalization (sQN) method with a search grid in-
terval k=[101 , 103 ] and ∆k=1.2 is typically 35 times slower than the simplified Q-normalization
(QN) method without scaling factor (see section 3.1.2). Note, that the sQN method offers the
advantage of a significantly higher accuracy compared to QN, as will especially be demonstrated
in chapter 6.
1
10
0
10
sQN
−1
10
tc [s]
−2
10
QN
−3
10
−4
10
−5
10 1 2 3 4 5
10 10 10 10 10
R
F IGURE 4.9.: Calculation time tc of QN and sQN algorithms depending on the number of bins
R. Test distribution (section 3.2.2): σRJ =0.05 UI, ADJ =0.2 UI (uniform DJ),
N =107 , K=50 evaluations.
With a bathtub curve that covers half of the unit interval (UI) and thus, consists of ≈150 k bath-
tub samples at Rsim =3.33·105 , sQN optimization of both tails takes several seconds. This may be
acceptable for system behavioral simulations where usually minutes or hours are spent to collect a
sufficient amount of data samples. With hardware jitter measurements or in production testing this
would be too time consuming, but here the bathtub is obtained using phase interpolators of coarse
time resolutions, that divide the UI into larger sized intervals. Assuming typically R=128 bins as
given with a 7 Bit phase interpolator, the optimization process is also several orders of magnitude
faster compared to the simulation.
36
4.2. P ERFORMANCE A NALYSIS
In this section the tail fitting performance of the proposed sQN method is analyzed. Therefore, test
distributions and error metrics from section 3.2 are used. According to the way of synthesizing
these distributions, the estimation error can be investigated with respect to varying distribution
shape (σRJ , ADJ , DJ type) and sample size N .
Accuracy of the extrapolated timing budget TJpp,est is evaluated by observing statistical spread
and bias of the estimation error over multiple evaluation runs. At least several hundred evaluations
are thus necessary to reliably judge the error behavior. Since it is not possible to predict whether
an implemented algorithm reveals convergence problems, statistical measures must be especially
robust against outliers. Therefore, median value Emed and interquartile range IQR are used,
according to the definitions from equation (3.17).
An initial error analysis is carried out by creating test distributions with different values of σRJ
and ADJ (uniform), where the median error bias of the tail fitting method is investigated. Results
are shown in figure 4.10(a). The obtained figure is symmetric and can thus be simplified by using
only the ratio σRJ /ADJ instead of both variables. In figure 4.10(b) a rotated view is obtained with
σRJ /ADJ as single dependent variable, which allows to discard σRJ . From a mathematical per-
spective the shape of a jitter distribution is fully described by this ratio, as long as the estimation
error is not influenced by the timing quantization of distributions. This is always the case for sim-
ulations where R, so that the estimation error Emed does not vary. In the example of figure 4.10
the default analysis configuration from table 4.1 has been used to construct the surfaces. Each
point on the surface is the median error value of K=250 evaluations where the estimated TJpp,est
values are obtained from fitted and extrapolated distribution tails. The true TJpp,true values for
error calculation are given by numerical approximations as described in section 3.2.2. Each of the
distributions uses N =108 samples with a resolution of Rsim =3.33·105 time divisions per UI.
In the following subsections an initial analysis investigates the influence of varying sample size
N . Then, the focus is put on the ratio σRJ /ADJ as distribution shape variable as well as on
different DJ types. Therefore, always the default values from table 4.1 are used for performance
evaluation, unless otherwise specified.
0.01
0.01
0.005
Emed
0.005
0 0
−3
10
−3 10
−3 2
10 10
−2 −2 −2 0
10 10 10 10
−1 −2
10 10
−1 0 −1 −4
σ 10 10 ADJ σRJ 10 10 σRJ/ADJ
RJ
(a) (b)
F IGURE 4.10.: Influence of both σRJ and ADJ on median estimation error Emed . With the
variable ratio σRJ /ADJ a rotated view is obtained where the symmetric repre-
sentation can be simplified and one of the two variables discarded.
37
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
0.62
0.6
0.58
pp,est
0.56
TJ
0.54
0.52
0.5
1k 5k 10k 50k 100k 500k 1M 5M 10M 50M 100M 500M 1G
F IGURE 4.11.: Example for extrapolation error over varying sample size N.
TJpp,true =0.523 UI, ADJ,uni =0.2 UI, σRJ =0.025 UI, K=250.
In figures 4.12(a) and 4.12(b) the influence of varying sample size is demonstrated with respect
to two different jitter ratios (σRJ /ADJ =1/8 and 1/128). The markers show median error values
obtained with two hundred realizations, while the dashed lines denote upper and lower quartiles as
statistical spread. With increasing sample size, both error bias as well as statistical spread decrease
toward zero, which empirically proves consistency of the scaled Q-normalization (sQN) approach.
0 0
10 10
−1 −1
10 10
Estimation error Emed
−2 −2
10 10
−3 −3
10 10
−4 −4
10 2 4 6 8 10
10 2 4 6 8 10
10 10 10 10 10 10 10 10 10 10
Sample size N Sample size N
(a) σRJ /ADJ =1/8, ADJ,uni =0.2 UI (b) σRJ /ADJ =1/128, ADJ,uni =0.2 UI
F IGURE 4.12.: Influence of varying sample size N on estimation error Emed , with two different
test distributions.
38
4.2. P ERFORMANCE A NALYSIS
The positive error bias is an effect caused by the Q-normalization principle. Bathtub functions
are transformed into Q-domain where they are represented in a linearized form. This linear behav-
ior of Q-tails is approached asymptotically, which introduces error bias for extrapolated tails. As
figure 4.13 demonstrates for a left Q-tail, fitted lines tend to overestimate the true timing budget,
especially at small sample size N . With a large number of jitter samples, bathtub curves can be
tracked down to deep probability levels. Simultaneously, also the asymptotic behavior of a Q-tail
is significantly reduced. Overestimated TJ values yield positive errors with pessimistic estimates,
which is a beneficial property of fitting methods based on Q-normalization. This general property
is valid for the sQN method as well. Figure 4.12 also highlights an additional influence of the
jitter ratio σRJ /ADJ or distribution shape on estimation performance. This effect is investigated
subsequently.
Exact
Q(x)
N << N >>
x x x
Estimated
F IGURE 4.13.: Asymptotic linearity of Q-tails as fundamental cause for error bias.
−1 −1
10 10
−2 −2
10 10
Emed
Emed
−3 −3
10 10
−4 −4
10 −2 −1 0
10 −2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
(a) (b)
N =108 N =107 N =106 N =105 N =104
F IGURE 4.14.: Influence of jitter ratio σRJ /ADJ,uni and sample size N on error.
39
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
−1 −1
10 10
−2 −2
10 10
Emed
Emed
−3 −3
10 10
−4 −4
10 −2 −1 0
10 −2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
(a) (b)
Sinusoidal Uniform Triangular Quadratic Curve
F IGURE 4.15.: Influence of jitter ratio σRJ /ADJ and DJ type on error. N =107 .
for the RJ dominant case, with a maximum around σRJ /ADJ,uni ≈1/4. This distribution shape
will later on also be utilized for worst case analysis.
So far, only uniform DJ has been considered for performance evaluation. Similar to the previous
plots, figures 4.15(a) and 4.15(b) thus demonstrate the estimation error and evaluation spread for
sinusoidal, uniform, triangular and quadratic curve shaped DJ. If the DJ type is changed to more
Gaussian-like shapes such as a triangle or a quadratic curve, estimates also degrade since the algo-
rithm tends to detect a single Gaussian-like peak instead of the steep tails at the distribution edges.
Only a very small percentage of the collected samples belongs to the true Gaussian tail, making a
correct tail detection very difficult. Therefore the estimation error becomes large, especially when
quadratic curve shaped DJ is combined with a small RJ component. Although this combination
is rather theoretical and unlikely to appear in real measurements, a possible way to handle the
problem is to use a specific parameter configuration optimized with respect to the desired working
region.
40
4.3. P ERFORMANCE O PTIMIZATION WITH D IFFERENT F ITNESS M EASURES
As was shown with the optimization scheme in section 4.1.1, the proposed sQN method yields
k-scaled Q-functions after the transform (see equation (4.3), figure 4.4(a)), and identifies scaling
factor k and fitting length n with optimum linearity. For this case the regression error becomes a
minimum, but the length of the linearized tail also approaches a maximum. Thus one may consider
a fitness measure based on the regression length of Q-tails as well. The criterion n̂ can for example
be defined as
nerr,min
n̂ = 1 − , nerr,min = n (4.14)
R σ̂err =min{σ̂err (n)}
where R is the number of bins per UI, here needed for normalization, and nerr,min is the length
which corresponds to the regression error minimum along a transformed Q-tail. n̂ maximizes the
regression length n or number of fitted tail samples over varying scaling factor k. That is, the
regression error σ̂err is still used for minimization along the first search dimension n, while n̂ is
applied along the second search dimension k.
In [117] Scholz provides a linearizing transform for generalized extreme value distributions,
and suggests a fitness criterion based on the ratio of regression error σ̂err and tail slope s:
T̂ = σ̂err /s (4.15)
Scholz analyzes the covariance structure of the regression error after performing the transform and
demonstrates, that the variance of the error depends on the slope s of the fitted line. Thus, T̂ can
be used to indicate the appropriateness of a fitted tail and to compensate the slope influence.
Equivalent to the regression length n̂, a length measure based on T̂ instead of σ̂err may be
defined. This measure also maximizes the regression length of fitted tails, and is denoted as n̂T .
nT,min
n̂T = 1 − , nT,min = n (4.16)
R T̂ =min{σ̂err (n)/s(n)}
n̂T also searches for the best suited scaling factor k and acts only along the second search dimen-
sion.
With the different fitness measures, now also the algorithmic implementation from figure 4.8
must be adapted. Therefore, in figure 4.16 the flow graphs of the fitting algorithms based on σ̂err
(left) and T̂ (right) optimization are plotted again. Depending on the selected type of algorithm,
either the regression length (n̂, n̂T ) or the regression error (σ̂err , T̂ ) is used for the outer loop with
scaling factor k. The nested loop remains the same for both cases. All four optimization criteria
are also summarized in table 4.2.
Estimation performance of the fitness measures in figure 4.17 is tested by carrying out K=250
evaluation runs over varying jitter ratio σRJ /ADJ . From figure 4.17(a) we notice that σ̂err pro-
vides a smaller error bias (expressed with markers as median values), while n̂ achieves a better
accuracy, or less statistical spread (expressed by the dashed lines as upper and lower quartiles).
An equivalent effect is observed with T̂ and n̂T . Although T̂ provides the smallest bias out of the
four fitness criteria, the lower quartiles evidence a large distance to the median values.
In order to determine an optimum trade-off between these two conflicting interests, a common
performance indicator must consider both error bias and dispersion. When recalling the perfor-
mance metrics from section 3.2, the estimation loss definition in equation (3.18) EL =|Emed | +
41
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
Qk =Q(CDF(x) · k) Qk =Q(CDF(x) · k)
n++ n++
reg. length reg. length
no Qk (n)≥0? no Qk (n)≥0?
yes yes
store as {n̂n (k), σ̂n (k)} store as{n̂T,n (k), T̂n (k)}
k=k·∆k k=k·∆k
scaling factor scaling factor
no k≥kmax ? no k≥kmax ?
yes yes
F IGURE 4.16.: Flow graphs of σ̂err based (left) and T̂ based (right) algorithmic principles.
1.5·IQR can be used. EL gives a simple confidence interval which is exceeded by less than 2.2%
of the evaluations for a normal error distribution.
With EL the proposed fitness measures can now be compared against each other as shown in
figure 4.18. From this representation we notice that the performance of pure regression error σ̂err
is not outperformed by T̂ due to its large statistical spread. Both n̂ as well as n̂T perform slightly
worse than σ̂err over a broad shape region. The reason why T̂ is outperformed by σ̂err , is the
large distance of lower quartiles from the median value. In fact, TJ estimates demonstrate skewed
distributions with heavy negative tails. A possible cause to this effect might be correlated errors
inside the regression model. In [117] Scholz de-correlates the error of regression parameters using
the inverse of the covariance matrix, but unfortunately this highly increases complexity of the
fitting algorithm and is thus impractical for computations.
Since the asymptotic tail behavior always guarantees for a positive error bias, one could argue
that the estimation loss can also be calculated using only the upper quartile distance. Negative
errors would thus not affect the worst case error, even if distributions are heavy tailed. Neverthe-
42
4.3. P ERFORMANCE O PTIMIZATION WITH D IFFERENT F ITNESS M EASURES
−1 −1
10 10
−2 −2
10 10
Emed
Emed
−3 −3
10 10
−4 −4
10 −2 −1 0
10 −2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
F IGURE 4.17.: Estimation performance of different fitness measures from table 4.2. Test distri-
butions are generated with the default analysis configuration from table 4.1.
0.03
0.02
EL
0.01
0 −2 −1 0
10 10 10
σRJ/ADJ
σ̂err n̂ T̂ n̂T
less, in order to avoid outliers and heavy tails as far as possible and to ensure a certain algorithmic
robustness, a Gaussian-like error behavior is preferred.
A veritable improvement of estimation performance can be achieved by combining fitness mea-
sures during the optimization procedure. So far, the whole optimization has been performed using
only one fitness measure out of the candidates from table 4.2. When recalling the flow graphs
in figure 4.16, the optimization process first starts with an initial search grid and then refines the
obtained optimum grid value down to the desired accuracy. This principle can also be used to
combine two different fitness measures. Out of the numerous possibilities two combinations are
investigated and presented subsequently.
In figure 4.19 the search grid resolution ∆k is varied using both σ̂err and n̂ as fitness measures.
For the logarithmically scaled search points, n̂ is first used to determine the initial rough grid
estimate, and the refinement is then continued with σ̂err in a local search environment. From the
resulting EL curves, the best results are obtained with a grid distance ∆k=1.2. A comparison
with the performance in figure 4.18 clearly shows the improvement.
43
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
0.02
0.015
EL
0.01
0.005
0 −2 −1 0
10 10 10
σRJ/ADJ
F IGURE 4.19.: Performance EL over varying search grid resolution ∆k={1.1, 1.2, 1.5, 1.8}.
The black curve is the reference from figure 4.18.
0.02
0.015
EL
0.01
0.005
0 −2 −1 0
10 10 10
σ /A
RJ DJ
F IGURE 4.20.: Performance EL with secondary search refinement ∆k={1.1, 1.2, 1.5, 1.8}. The
black curve is the reference from figure 4.18
In figure 4.20, after complete optimization with n̂, a second refinement step with σ̂err is addi-
tionally performed in a local environment around the first estimate. The environment is specified
with the same scaling factor ∆k. The performance results are similar to those before and thus,
again better than simple σ̂err optimization. Considering the increased optimization effort for this
scenario the first one is the more convenient choice.
Summarizing these results, a well suited combination of n̂ and σ̂err measures is obtained, when
a search grid resolution ∆k=1.2 first provides an initial estimate with n̂, which is then refined by
σ̂err . Subsequently this combined optimization scenario is referred to as ĉ1.2 .
Note that the error maxima in figures 4.18 to 4.20 are always located at the same jitter ratio
σRJ /ADJ ≈1/4. This means, the worst case shape is constant and independent from the inves-
tigated fitness measures, and can thus be used for simplified worst case analysis. The presented
analyses used only test distributions of uniform type DJ and thus, the optimized ĉ1.2 scenario is
44
4.3. P ERFORMANCE O PTIMIZATION WITH D IFFERENT F ITNESS M EASURES
theoretically only valid for this single DJ shape. In other cases one would thus also have to search
for other combinations as well. However, the uniform distribution is here considered as a good
trade-off between easily separable sinusoidal and difficult Gaussian-like shapes. This allows to
use ĉ1.2 as algorithmic scenario for all DJ shapes.
45
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
see how an adequate choice improves fitting performance. The definition of estimation loss EL
from equation (3.18) is therefore reused as performance indicator. To investigate the influence of
outliers caused by the algorithm, further the kurtosis κ is utilized as a presence-of-outlier indicator
which has to be minimized. Subsequent analyses are carried out with the default parameters from
table 4.1.
The parameter ∆Tt specifies a constant part of the unit interval (UI), and hence, it defines a
minimum number of tail samples nmin which must be included with the linear regression analysis.
For a simulator with pre-specified time resolution Rsim , nmin is given by:
∆Tt has to be chosen carefully in order to afford correct TJ estimation. If selected too large,
bathtub samples that do not belong to the Gaussian tail are used for tail fitting, which results in a
high estimation error. If selected too small, outliers may be caused by wrong convergences of the
linear regression stage, simply because the initial tail region is only supported by a few bathtub
samples. Especially RJ dominant distributions consist of highly varying tail endings and thus,
suffer from this effect.
Median error Emed , interquartile range IQR, performance EL and kurtosis κ, depending on
both ∆Tt and Pt,min are investigated in the rows of figure 4.22. The jitter ratio σRJ /ADJ,uni =1/4
corresponds to the worst case shape from the performance analysis in figure 4.18, where the esti-
mation loss is maximum. The other analysis parameters are set to the default configuration from
table 4.1. Three different optimization criteria with σ̂err , n̂ and ĉ1.2 from section 4.3.1 are investi-
gated in the different columns.
The plots in the first row show the median error Emed , which is approximately constant for a
broad range of ∆Tt values and decreases toward higher values of Pt,min , until a critical Pt,min
value is reached. Here the results become highly unstable unless the selected ∆Tt is close to the
optimum Gaussian tail length of the given test distribution, which is ∆Tt ≈0.1UI. For larger values
of ∆Tt the median error again increases.
The interquartile range IQR in the subfigures of the second row remains constant for a large
range of ∆Tt and reaches a minimum at the correct Gaussian tail length. The IQR increases
toward higher Pt,min values, which confirms the importance of bathtub samples at lowest prob-
ability levels. In fact, the tail fitting method looses accuracy when these samples are cut off by
the probability threshold Pt,min . The interquartile range also highlights a large influence on the
estimation loss EL , which has been calculated according to the definition from equation (3.18),
and is depicted in the third row of figure 4.22.
In the bottom row, the kurtosis surfaces display the outlier behavior of the scaled Q-normali-
zation method. Again, optimum kurtosis κ is only achieved with a known tail length. As soon as
the probability threshold discards parts of the measured tail (Pt,min >0), κ increases significantly,
and thus, outliers appear.
Considering that both EL and kurtosis have to be minimized, the best performance for all three
fitness measures is achieved with Pt,min =0 and ∆Tt ≈0.1 UI. Note that this parameter configura-
tion can only be used for the given test distribution but not for other shapes. This means, a priori
knowledge of the ideal tail length is needed in order to obtain optimum fitting results.
When comparing the three fitness measures σ̂err , n̂ and ĉ1.2 , the least statistical spread is
achieved with n̂ at Pt,min =0 (figure 4.22(e)). However, the overall estimation loss EL behaves
significantly better with ĉ1.2 and thus, achieves best overall performance with ĉ1.2 optimization.
46
4.3. P ERFORMANCE O PTIMIZATION WITH D IFFERENT F ITNESS M EASURES
Emed
Emed
E
−1 −1 −1
Pt,min
−7 10 Pt,min
−7 10 Pt,min
−7 10
10 ∆ Tt 10 ∆ Tt 10 ∆ Tt
−1 −1 −1
10 10 10
−2 −2 −2
IQR
IQR
IQR
10 10 10
−3 −3 −3
10 10 10
−3 −3 −3
10 10 10
−5 −5 −5
10 10 10
−5 −5 −5
10 10
−3 10 10
−3 10 10
−3
−1 −1 −1
−7 10 −7 10 −7 10
Pt,min 10 ∆T Pt,min 10 ∆T Pt,min 10 ∆T
t t t
EL
EL
E
−1 −1 −1
Pt,min
−7 10 Pt,min
−7 10 Pt,min
−7 10
10 ∆ Tt 10 ∆ Tt 10 ∆ Tt
2 2 2
10 10 10
Kurtosis
Kurtosis
Kurtosis
1 1 1
10 10 10
0 0 0
10 10 10
−3 −3 −3
10 10 10
−5 −5 −5
10 10 10
−5 −5 −5
10 −3
10 10 −3
10 10 −3
10
−1 −1 −1
−7 10 −7 10 −7 10
Pt,min 10 ∆ Tt Pt,min 10 ∆ Tt Pt,min 10 ∆ Tt
F IGURE 4.22.: Emed (top), IQR, EL , and kurtosis κ (bottom) for ∆Tt versus Pt,min parameter
surfaces. Three optimization criteria σ̂err (left), n̂ (middle) and ĉ1.2 (right) are in-
vestigated. Test distribution: σRJ =0.05 UI, ADJ,uni =0.2 UI, N =107 , K=250.
∆Pt defines a probability region or interval where bathtub samples at lowest probability levels
must be included for regression analysis. Distributions with a small RJ component yield steep
bathtub tails which are tracked by few tail samples, while for the RJ dominant case many tail
samples are needed. However, for both cases the Gaussian probability region is approximately
constant. The parameter ∆Pt instead of ∆Tt thus, offers a much more robust way of initial tail
selection, without prior knowledge of the distribution shape. Further, random variations at the
47
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
Emed
Emed
E
−1 −1 −1
10 10 10
−2 −2 −2
IQR
IQR
IQR
10 10 10
−3 −3 −3
10 10 10
−3 −3 −3
10 10 10
1 1 1
−5 10 −5 10 −5 10
10 3 10 3 10 3
10 10 10
5 5 5
−7 10 −7 10 −7 10
Pt,min 10 ∆ Pt Pt,min 10 ∆ Pt Pt,min 10 ∆ Pt
EL
E
2 2 2
10 10 10
Kurtosis
Kurtosis
Kurtosis
1 1 1
10 10 10
0 0 0
10 10 10
−3 −3 −3
10 10 10
1 1 1
−5 10 −5 10 −5 10
10 3 10 3 10 3
10 10 10
5 5 5
−7 10 −7 10 −7 10
Pt,min 10 ∆ Pt Pt,min 10 ∆ Pt Pt,min 10 ∆ Pt
F IGURE 4.23.: Emed (top), IQR, EL and κ (bottom) for ∆Pt versus Pt,min parameter surfaces.
Three optimization criteria σ̂err (left), n̂ (middle) and ĉ1.2 (right) are investigated.
Test distribution: σRJ =0.05 UI, ADJ,uni =0.2 UI, N =107 , K=250.
outermost tail endings, which are caused by the probability granularity 1/N , can directly be faced
by choosing ∆Pt sufficiently large.
In figure 4.23 again the fitting behavior of the scaled Q-normalization method is analyzed, this
time as a function of the parameters ∆Pt and Pt,min . The selected test distribution and perfor-
mance measures are the same as in figure 4.22.
For all three optimization criteria σ̂err , n̂ and ĉ1.2 a convergence limit for the variable product
∆Pt ·Pt,min =const is noticed. It is given by the Gaussian tail part of the test distribution. An
48
4.3. P ERFORMANCE O PTIMIZATION WITH D IFFERENT F ITNESS M EASURES
upper probability level Pup for tail selection can be defined according to this product:
Pup must belong to the Gaussian part of the distribution tail in order to guarantee for correct fitting,
and hence, it must be significantly smaller than the minimum Gaussian tail amplitude At,min .
Since a Q-tail approaches the linear behavior asymptotically, such an exact probability level cannot
be determined but has to be approximated, as will be demonstrated later in section 5.2. When Pup
is located in the convergence region well below the tail amplitude At,min , a similar behavior of
Emed , IRQ and EL is noticed as with the analysis of ∆Tt before. Emed basically decreases
toward higher values of Pt,min , but the interquartile range IQR and also the overall estimation
loss EL increase. For all three optimization criteria σ̂err , n̂ and ĉ1.2 a minimum estimation loss
EL and smallest kurtosis κ is obtained, if also the threshold parameter Pt,min is at its minimum.
Estimation performance depending on the parameters ∆Pt and Pt,min must also be investigated
with respect to varying distribution shape σRJ /ADJ . The resulting estimation loss surfaces for
σRJ /ADJ =1/256 are plotted in figure 4.24 and demonstrate a behavior which is similar to the
prior figures. Besides less estimation error in the convergence region which is due to a better
suited distribution shape, the convergence limit has now also moved toward a lower ∆Pt ·Pt,min
variable product because of the smaller Gaussian tail amplitude.
EL
EL
F IGURE 4.24.: EL for parameter surfaces ∆Pt versus Pt,min , with σ̂err (a), n̂ (b) and ĉ1.2 (c)
optimization criterion. Test distribution: σRJ /ADJ =1/256, ADJ,uni =0.2,
σRJ =0.00078125, N =107 , K=250.
This shows, that it is not possible to specify an ideal parameter configuration which yields best
fitting performance for the general case of arbitrary distribution shapes. Instead, one has to rely on
a suboptimal configuration, which guarantees for correct algorithmic convergence and minimizes
the influence on estimation performance. From the EL surfaces in figures 4.23 and 4.24 at least
∆Pt ≥102 is recommended to afford correct tail fitting. Further, without Pt,min the kurtosis of
estimates demonstrates a minimum outlier presence, combined with a maximum selection range
for ∆Pt . Thus Pt,min can be omitted and replaced with the minimum probability granularity
Pt,min =1/N instead. Equation (4.18) is thus reformulated with the condition:
As long as the upper probability level for tail selection Pup is located significantly below the
minimum Gaussian tail amplitude At,min , jitter distributions will be fitted correctly. If ∆Pt is
chosen too large, fitted tails are forced to include samples from the DJ component, which misleads
the extrapolation result. With a given minimum amplitude At,min it is thus possible to define
a suitable parameter range for ∆Pt . As a general selection criterion, ∆Pt should be chosen as
large as possible, but without violating equation (4.19). In section 5.2 design aspects for jitter
49
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
analysis systems will be investigated, and analyses will focus on ideal tail selection with more
detail, especially on a relation between At,min and the test distribution shape.
To summarize these results, best tail fitting behavior for the scaled Q-normalization (sQN)
method is obtained, when using the fitness criterion ĉ1.2 , together with ∆Pt as conservative pa-
rameter for tail selection. The other two parameters ∆Tt and Pt,min can be discarded. A suited
selection range for ∆Pt is given with:
Further, ∆Pt should always be as large as possible, which yields the default values
(
103 if N ≥ 106
∆Pt = (4.21)
N/103 if N ≤ 106
These default values for ∆Pt are chosen such that the upper tail selection bound ∆Pt /N is lo-
cated significantly far above the probability granularity 1/N of collected jitter distributions and
thus, avoids misleading outliers. Even if distribution tails are very flat (RJ dominant case) and
suffer from statistical noise, ∆Pt =103 guarantees for correct convergence if N ≥106 . However
at small sample sizes, ∆Pt must also be selected smaller, in order to fulfill condition (4.19). The
minimum Gaussian amplitude At,min which can be fitted correctly by the sQN method, thus, must
be significantly larger than
At,min 10−3 (4.22)
The investigated test distributions from section 3.2.2 violate this amplitude minimum only for
extreme cases with triangular and quadratic curve shaped DJ. Further, for N =104 we only have
∆Pt =101 and thus, partly accept outliers. The default ∆Pt configuration is applied subsequently.
−1
10
−1
10
−2
10
Emed
EL
−2
10
−3
10
−4 −3
10 −2 −1 0
10 −2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
(a) (b)
N =108 N =107 N =106 N =105 N =104
F IGURE 4.25.: Emed and EL performance of ĉ1.2 based optimization over varying σRJ /ADJ,uni
and sample size N . N ={104 , 105 , 106 , 107 , 108 }.
50
4.4. P ERFORMANCE O PTIMIZATION WITH Q-D OMAIN T HRESHOLD
−1
10
−1
10
−2
10
Emed
EL
−2
10
−3
10
−4 −3
10 −2 −1 0
10 −2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
(a) (b)
Sinusoidal Uniform Triangular Quadratic Curve
F IGURE 4.26.: Emed and EL performance of ĉ1.2 based optimization over varying σRJ /ADJ
and DJ shape: sinusoidal, uniform, triangular and quadratic curve. N =107 .
afford evaluations without causing too many outliers. As figure 4.25 shows, the median estimation
error Emed with optimized parameters usually performs slightly worse than with prior tail search
(expressed by the dotted lines), but the estimation loss EL is improved due to less statistical spread
combined with complete outlier suppression. In fact, the kurtosis (not shown in the figures) of the
optimized search criterion has become significantly smaller.
If again the DJ shape is varied, as depicted in figure 4.26, a similar performance improvement
for EL is noticed. With quadratic curve shaped DJ an error peaking in the lowest σRJ /ADJ
region is additionally noticed, which is due to wrong tail fits. In this region, bathtub tails become
extremely steep and At,min reaches below the bound given by (4.19). The fitting algorithm is thus
not able to detect the correct tail amplitude with the given bathtub samples anymore, and rather
converges toward the Gaussian-like DJ shape instead of the steep RJ component.
51
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
2
k=16
k=4
k=64
0
Q k=1
Quantile [Q]
min
−2
−4
In Q-domain, functions are described in terms of the standardized variable q=(x − µ)/σ (3.6),
which leads to a simple regression model for Gaussian tails represented as straight line. The
variable q thus also indicates the number of standard deviations one moves away from the mean
value of the Gaussian model. This means, the Qmin threshold defines the tail part as distance from
the Gaussian model mean in terms of its standard deviation. As a direct consequence, the fitting
region is automatically adjusted according to the expected Gaussian model.
The example in figure 4.27 uses a Q-domain threshold of Qmin =−1 equal to one standard devi-
ation. This value corresponds to the lower inflection point of a Gaussian PDF, located at x=µ−1σ.
When searching for the best suited regression line over varying k, it is thus assumed that measured
jitter distributions follow a Gaussian function at least up to the inflection point. In scaled Q-domain
the search can thus be carried out without knowing the Gaussian model parameters. For k=1 the
measured Q-tail already starts to bend before reaching the q=−1 level, indicating that the tail
probabilities increase faster than the expected Gaussian with amplitude A=1/k=1. Equivalently,
also for k=64 the Q-tail does not follow a linear behavior. The optimized tail amplitude is found
at k=7.15 where the linear part even reaches beyond the q=0 level, which means that more than
half of the Gaussian function can be fitted into the measured tail.
Obviously, different values for the threshold parameter Qmin are possible. An analysis of the
suited parameter range is thus needed and carried out subsequently. Different optimization sce-
narios and fitness measures also require a performance comparison, in order to identify a best
suited configuration. Similar to the ĉ1.2 criterion with the parameter ∆Pt (section 4.3.2), the
Qmin threshold also defines a minimum amplitude At,min and thus, limits the Gaussian tail model
search. An according functional relation will be derived in section 5.2. As a beneficial property,
Qmin allows for the derivation of an exact equation compared to the rough approximation for ∆Pt
in equation (4.19).
52
4.4. P ERFORMANCE O PTIMIZATION WITH Q-D OMAIN T HRESHOLD
Qk =Q(CDF(x) · k) Qk =Q(CDF(x) · k)
n++ n++
Qk (n)≥Qmin ? Qk (n)≥Qmin ?
no no
yes yes
Qk (n)≥0?
no
yes
k=k · ∆k k=k · ∆k
scaling factor scaling factor
no k≥kmax ? no k≥kmax ?
yes yes
σ̂n,k =min{σ̂n (k)} σ̂n,k =min{σ̂n (k)}
... ...
F IGURE 4.28.: Flow graphs for algorithmic principles based on Q-domain threshold, with mini-
mum Qth,min (left) and constant Qth,c (right) threshold scenario.
also allows for any larger tail lengths. Equivalent to the optimization principles from section 4.3
the best suited tail part is then identified according to the smallest fitness value along the tail.
The flow graphs for these two algorithmic principles are plotted in figure 4.28 and subsequently
denoted as Qth,c for constant and Qth,min for minimum threshold analysis. The last blocks have
been omitted since they are equivalent to prior algorithms (see figure 4.16). The left graph de-
scribes the Qth,min based fitting principle, where Qmin is now included into the algorithmic struc-
ture. As long as the Q-tail values Qk (n) are located in the tail region, and thus Qk (n)<Qmin
(also see figure 4.27), they will recursively be added by the algorithm but not used for minimum
search. As soon as the Q-domain threshold is reached, the minimum search finds the best suited
tail length inside the interval Qmin <Qk (n)<0. This minimum value has to be identified for every
scaling factor k in order to allow for Gaussian amplitude search.
The right graph in figure 4.28 describes the simple Qth,c scenario, where only the initial tail part
up to Qmin is used for regression analysis. Other tail lengths are not allowed. Note, that both flow
graphs only describe the search with an initial coarse search grid, and thus have to be followed by
a refinement stage in order to yield the final results. The refinement stages operate with the same
optimization criterion as for the respective search grid.
53
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
2
0.015 0.03 10
0.01 0.02
Kurtosis
Emed
EL
1
10
0.005 0.01
0 0 −2 −1 0 −3 −2 −1 0
−2 −1 0
10 10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
(a) {Qth,min , σ̂err }, Emed (b) {Qth,min , σ̂err }, EL (c) {Qth,min , σ̂err }, κ
2
0.015 0.03 10
0.01 0.02
Kurtosis
Emed
EL
1
10
0.005 0.01
0 0 −2 −1 0 −3 −2 −1 0
−2 −1 0
10 10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
0.01 0.02
Kurtosis
Emed
EL
1
10
0.005 0.01
0 0 −2 −1 0 −3 −2 −1 0
−2 −1 0
10 10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
(g) {Qth,c , σ̂err }, Emed (h) {Qth,c , σ̂err }, EL (i) {Qth,c , σ̂err }, κ
2
0.015 0.03 10
0.01 0.02
Kurtosis
Emed
EL
1
10
0.005 0.01
0 0 −2 −1 0 −3 −2 −1 0
−2 −1 0
10 10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
F IGURE 4.29.: Emed (left), EL (middle) and kurtosis κ (right) over varying distribution shape
σRJ /ADJ . In the different rows four algorithmic principles with combinations of
{Qth,min , Qth,c } and {σ̂err , T̂ } are investigated. The Qmin threshold is chosen
as −Qmin ={1.0, 1.2, 1.4, 1.6, 1.8, 2.0}, uniform DJ type, N =107 , K=250.
54
4.4. P ERFORMANCE O PTIMIZATION WITH Q-D OMAIN T HRESHOLD
Both flow graphs can either use σ̂err or T̂ =σ̂err /s as fitness measures to drive the minimum
search. Together with the two algorithmic principles Qth,c and Qth,min , thus four combinations
are obtained which can be compared against each other. Other fitness measures such as n̂ or n̂T
cannot be applied, since Qmin also affects the regression length over varying scaling factor k.
In figure 4.29 the estimation performance for each of the four algorithm combinations is de-
picted. The different subfigure rows correspond to these combinations, while the columns de-
scribe the respective behavior of median error Emed , estimation loss EL and kurtosis κ over
varying distribution shape σRJ /ADJ . While EL describes the estimation performance, κ indi-
cates the presence of outliers, as already described in section 4.3.2. Positive peaks in the course
of κ indicate strong outliers and have to be avoided as far as possible. Further, each subfig-
ure also investigates the influence of Qmin parameter variations by showing different curves for
−Qmin ={1.0, . . . , 2.0}.
From the Emed curves in the left column of figures 4.29 (a,d,g,j) we notice that the Qth,c sce-
nario has a better influence on error bias compared to Qth,min . With decreasing threshold Qmin ,
error bias can be reduced significantly. Unfortunately, this benefit is payed with a higher statis-
tical spread, as can be seen from the EL plots in 4.29 (b,e,h,k) where the common performance
is changed only marginally. Considering the worst case distribution shape at σRJ /ADJ =1/4,
in fact the optimum compromise between error bias and spread can be found in the interval
Qmin ∈[−1.2, −1.0].
Additionally an increase of kurtosis is observed at largest shape values for Qmin ≤−1.4. This
indicates a high uncertainty of the algorithm when choosing a small outermost fitting region. In
fact, with RJ dominant shapes the Q-tails become linear over a wide probability range, and thus,
the algorithm suffers from random noise at the outer tails.
Comparing the two fitness measures σ̂err and T̂ , the first one tends to produce estimates that
are less affected by outliers. This can especially be observed when comparing the kurtosis of
figures 4.29(c) and 4.29(i), with 4.29(f) and 4.29(l). Hence, the Qth,c optimization scenario in
combination with σ̂err as fitness measure offers the best choice out of the four candidates.
In order to get a better impression on how Qmin affects estimation performance, in figure 4.30
the four algorithm candidates are compared with respect to varying −Qmin ={0.5, . . . , 2.4}. The
selected test distribution is σRJ /ADJ =1/4 to realize the worst case shape. A visible difference in
performance appears as soon as Qmin ≤−1.0. The Qth,c optimization estimates the true TJ value
slightly better than Qth,min , which can be observed in both Emed and EL figures. The kurtosis
of all four algorithms follows a flat course, until at some point a high peak is observed. This
peak is negative skewed, indicating that large negative outliers are present in the distribution of TJ
estimates.
A closer analysis of these outliers highlights, that the regression stage has converged to the
outermost tail part by fitting a Gaussian model of very small amplitude. Due to the small Qmin
threshold parameter, the algorithm was in fact forced to carry out the regression analysis at the
outermost tail part of distributions, which is highly affected by statistical random variations. The
further Qmin is moved toward the outer tail, the higher the risk that outliers will occur. The only
way to avoid such a high risk is thus to increase Qmin .
In order to determine an optimum value of Qmin which can be used for various distribution
shapes, a similar analysis with respect to different DJ types is carried out. Figure 4.31 shows the
performance behavior of Qth,c scenario with σ̂err fitness measure and different DJ types. A good
compromise is found at Qmin =−1.2. This value is located in the minimum region of estimation
loss curves 4.31(c), and provides correct tail fitting results at an acceptable risk, as can be seen
from the kurtosis in 4.31(e). However, the quadratic curve DJ produces misleading outliers even
for Qmin ≥−1.2. The reason is the same as described before. The Q-tails of test distributions have
become very steep, so that random variations again easily cause outliers.
55
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
0.01
Emed
IQR
EL
0.01 0.02
0.005
0.005 0 0.01
0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5
−Qmin −Qmin −Qmin
0
Skewness
Kurtosis
−1
1
10
−2
(d) (e)
Qth,c , T̂ Qth,c , σ̂err Qth,min , T̂ Qth,min , σ̂err
F IGURE 4.30.: Emed (a), IQR (b), EL (c), skewness (d) and kurtosis (e) of the four algorithmic
principles from 4.29 plotted over −Qmin ={0.5, . . . 2.4}. Test distribution (worst
case shape): σRJ /ADJ =1/4, uniform DJ, N =107 , K=250.
0.03 0.04
0.01
med
IQR
EL
0.02 0.03
E
0.005
0.01 0.02
0 0 0.01
0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5
−Qmin −Qmin −Qmin
0
Skewness
Kurtosis
−1
1
10
−2
(d) (e)
Uniform Sinusoidal Triangular Quadratic Curve
F IGURE 4.31.: Emed (a), IQR (b), EL (c), skewness (d) and kurtosis (e) of DJ types over
−Qmin ={0.5, . . . , 2.4}, Qth,c , σ̂err scenario. Test distributions (worst case):
σRJ /ADJ =1/4 (uni.), 1/2 (sin.), 1/8 (tri.), 1/16 (quad.). N =107 , K=250.
56
4.4. P ERFORMANCE O PTIMIZATION WITH Q-D OMAIN T HRESHOLD
−1
10
−1
10
−2
10
Emed
EL
−2
10
−3
10
−4 −3
10 −2 −1 0
10 −2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
(a) (b)
N =108 N =107 N =106 N =105 N =104
F IGURE 4.32.: Emed and EL performance for Qth,c criterion with Qmin =−1.2 over varying
σRJ /ADJ,uni and sample size N . N ={104 , 105 , 106 , 107 , 108 }.
−1
10
−1
10
−2
10
Emed
EL
−2
10
−3
10
−4 −3
10 −2 −1 0
10 −2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
(a) (b)
Sinusoidal Uniform Triangular Quadratic Curve
F IGURE 4.33.: Emed and EL performance for Qth,c criterion with Qmin =−1.2 over varying
σRJ /ADJ and DJ type: sin., uni., tri. and quad. curve. N =107 .
Similar to section 4.3, the optimized Q-domain threshold method can also be compared with the
original simple σ̂err optimization from section 4.2. According to the previous results, the constant
Q-domain criterion Qth,c uses a threshold of Qmin =−1.2. Figure 4.32 shows, that the median
estimation error Emed of the threshold method usually performs slightly worse when compared
with the original tail fitting method (dotted lines). However, the overall estimation loss EL is
better for N ≥107 , and comparable to the ĉ1.2 based algorithm from section 4.3. For N ≤106
the estimation loss EL is larger. The Q-domain threshold also improves the performance of all
investigated DJ types, as can be seen in figure 4.33.
The error peaking in the lowest σRJ /ADJ region, which could be observed with the optimized
ĉ1.2 method at quadratic curve shaped DJ (figure 4.26), is not present anymore. The observable
minimum tail amplitude At,min is now given by the search range limit for kmax =1/At,min . How-
ever, the Qmin threshold method is also less robust against outliers.
57
4. A FAST AND ACCURATE J ITTER A NALYSIS M ETHOD
4.5. Summary
A complete approach to Gaussian tail fitting, referred to as scaled Q-normalization (sQN) was
proposed and realized. Therefore, the quantile normalization principle was embedded into a three
dimensional optimization scheme for Gaussian model parameter search with unknown tail am-
plitude A, mean µ and standard deviation σ. Due to efficient recursions (equations (4.5)), this
optimization can be performed very fast. The calculation time linearly depends on the selected
number of bins R for a jitter distribution, and was investigated in figure 4.9. The algorithm was
implemented with C/C++ and yields very accurate TJpp estimates as the performance analysis
in section 4.2 demonstrates. The resulting estimation error bias is always positive and thus pes-
simistic, which is a further beneficial property of algorithms based on quantile normalization.
Causes for performance degradation of the method can be summarized as follows:
• A small sample size N of collected distributions, as shown in figures 4.14, 4.25 and 4.32.
• The fitting algorithm operates on distribution shapes in the worst case region, for example
σRJ /ADJ,uni ≈1/4 at uniform type DJ.
• The DJ shape is Gaussian-like but bounded, as shown in figures 4.15, 4.26 and 4.33.
Especially the last point leads to jitter distributions where only a very small percentage of the
collected samples belongs to the Gaussian tail part. The only way to deal with this problem is to
increase the sample size N and thus, to prolongate the Gaussian tail, but as tail samples become
rarely with larger jitter amplitude, acquisition time also increases exponentially.
In order to improve the estimation performance as well as the robustness of the sQN method, two
algorithmic approaches have been investigated. The first one, ĉ1.2 in section 4.3, uses conservative
fitting parameters in probability domain. It suggests an initial fitting region ∆Pt ≥102 which
covers at least two decades, in order to avoid outliers caused by statistical tail variations. From
this principle, a combination which utilizes both regression length and error, has been identified
as best suited optimization criterion. According to equation (4.19), the minimum tail amplitude
At,min must be located significantly above the minimum tail fitting region given by ∆Pt /N .
The second approach, Qth,c in section 4.4, uses a constant threshold Qmin in scaled Q-domain,
which defines the Gaussian tail region in terms of standard deviations beside the model mean.
This representation form allows for a flexible choice of the tail part. With a smaller parameter
Qmin , estimation accuracy is improved, but the risk for outlier occurrence is also increased. As an
acceptable compromise Qmin =−1.2 has been identified.
Both approaches improve the estimation performance compared to the simple algorithm based
on regression error σ̂err from section 4.2. For uniform DJ, N =106 and worst case test distribu-
tions, the error bias is <2% and the overall error is generally <3% in more than 97.5% of the cases.
Although the Q-domain threshold scenario seems to slightly outperform the ĉ1.2 scenario with dif-
ferent DJ types, it clearly performs worse at smaller sample sizes N ≤106 (see figures 4.25(b)
and 4.32(b)). In the subsequent chapter this effect is shown to originate from a larger error varia-
tion of Qth,c if only the outermost tail part can be used for a fit. Further, a relation between sample
size N , minimum tail amplitude At,min and threshold Qmin was derived. This also includes a
more detailed comparison of both algorithms.
So far, the proposed sQN method with its two algorithmic versions has only been applied to
simulated jitter distributions, where the simulator timing resolution Rsim is chosen sufficiently
high, so that an influence on estimation performance can be neglected. A coarse time quantization
of distributions introduces a limited, discrete amount of bins. This is an inherent property of hard-
ware systems and has to be investigated thoroughly to allow for a hardware implementation of the
proposed method. This issue will especially be addressed in the subsequent chapter. Summarized
parts of this chapter have also been published in [C1,C8,C9].
58
5. Hardware Design Aspects
This chapter focuses on hardware related design aspects for the scaled Q-normalization (sQN)
method. So far, the proposed approach has only been considered as a pure mathematical optimiza-
tion scheme, which is applied to simulated distributions or behavioral models to investigate the
impact of timing jitter on system performance. As an important application, one also likes to use
the method together with real jitter data collected in measurements. However, due to the limited
precision, such devices introduce certain quantization effects and include additional error sources
the analysis method has to cope with.
The subsequent sections investigate these effects, and try to derive empirical relations or de-
scribe the resulting error behavior and limitations. The idea is to give the designer useful hints on
how to select key parameters for a jitter measurement system, in order to guarantee a certain target
accuracy and robustness. Thus, a complete design guideline for the sQN method from chapter 4 is
provided.
A key advantage of jitter distributions is that they can be collected on-chip using built-in jitter
measurement (BIJM) systems. With data frequencies in the GHz range, off-chip instrumentation
may include significant noise caused by interconnect wires, which can easily affect measurements.
In such cases, the impact of external noise can be eliminated with a direct on-chip measurement.
Hence, in recent years a broad variety of BIJM topologies and implementations [16–18, 35, 48,
57, 60, 65, 66, 79, 100, 129, 131] has been realized as design-for-test (DFT) structures to support
production tests as well as on-chip diagnostics [12–14,86,130]. Although subsequent analyses for
the scaled Q-normalization method are basically valid for both instrumentation devices and BIJM
systems, the focus of this chapter is more on future oriented on-chip applications.
The timing jitter of a PLL is analyzed by measuring IO jitter, defined as the time difference
between reference frequency and PLL output clock (section 2.1.2). In the case of serial high-
speed interfaces with a clock and data recovery (CDR) unit, the reference frequency is given by
the analog input data, and jitter values are measured between bit transitions of the input signal
and the PLL output clock. In order to correctly quantify IO jitter, thus, an exact time interval
measurement has to be performed. Such measurements are typically realized with an adjustable
delay element inside the PLL under test, as depicted in figure 5.1. A binary phase detector (PD)
compares the delayed output clock against the transition edges of the serial data stream. If the
clock precedes the data edge, a logical one is created at the output, otherwise a zero. After N
clock cycles, the counter state reflects the jitter probability for the selected delay value. One can
sequentially step the delay over the whole bit period and thus, measure the complete probability
function of IO jitter. The smallest step size defines the achievable number of bins R in a unit
interval (UI), or time resolution 1/R.
A significant speed-up of the measurement is achieved if up to R PDs and counters are used in
parallel, together with a delay line [16,66]. Further, the PD can be replaced by a simple D flip-flop
or data latch if the input data is known. That is, the recovered data from the flip-flop is compared
against the expected original data, which again allows for detecting errors. However, the simple
BIJM principle in figure 5.1 measures only one bathtub point at a time. With N received bits
the probability value for the selected delay is tracked down to the BER level 1/N , which also
forms the probability granularity of the measurement. A complete bathtub measurement must be
59
5. H ARDWARE D ESIGN A SPECTS
Input Recovered
Data PLL
Clock
Delay
Counter
PD Jitter
Analysis
PLL Under Test
60
5.1. TAIL PARAMETERS OF T EST D ISTRIBUTIONS
0
1.2 1 10
0.8 sin
1.15 tri qua sin −1
10
qua
0.6
2⋅µt / ADJ
σt / σRJ
At
1.1 tri uni
uni 0.4
−2
uni 10 qua
1.05 tri
sin
0.2
−3
1 0 10 −2 −1 0
−2 −1 0 −2 −1 0
10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
F IGURE 5.2.: Fitted test distributions: tail parameters standard deviation σt (a), mean µt (b), and
amplitude At (c) over varying shape σRJ /ADJ , N =108 , R=Rsim =3.3·105 .
A possible way to deal with this problem is a fast numerical tail approximation that reflects
the realistic behavior of the sQN method. Therefore, ideal test distributions are created by con-
volving RJ and DJ components, and bathtub tails are simulated down to the target BER=10−12 ,
so that the noise of random variations is suppressed. Then quantization effects of N and R are
re-introduced. The applied fitting method thus yields the same tail parameters which otherwise
have to be estimated from median values of time consuming, statistical evaluations. This allows
to quickly characterize the average fitting behavior of the method. By applying this principle, in
the subsequent sections thus, empirical relations for tail parameters are derived.
61
5. H ARDWARE D ESIGN A SPECTS
0 0
10 10
−1 −1
10 10
Amed
Amed
−2 −2
10 10
−3 −3
10 −3 −2 −1 0
10 −3 −2 −1 0
10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
F IGURE 5.3.: Comparing the median of tail amplitude At over varying jitter ratio σRJ /ADJ with
fast numerical approximations (dashed lines). N =108 , K=250, ∆Pt =105 .
TABLE 5.1.: Coefficients for equations (5.3) and (5.4), with sQN ĉ1.2 algorithm, R=Rsim .
The four curves in figure 5.2(c) follow a linear behavior over a large range of jitter ratios,
until the tail amplitude is close to the maximum value At =1. An empirical relation between tail
amplitude and jitter ratio can thus be obtained using simple linear functions. For the resulting
regression coefficients a logarithmic scaling of both axes must additionally be considered, thus:
62
5.1. TAIL PARAMETERS OF T EST D ISTRIBUTIONS
2
2 10
uni
1.8
qua
(σt/2µt) / (σRJ/ADJ)
σt / σRJ 1.6
tri
1
10
1.4 tri
qua
uni sin
1.2
sin
0
1 −3 −2 −1 0
10 −2 −1 0
10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
(a) (b)
σt σt /2µt
F IGURE 5.4.: Normalized standard deviation σRJ (a) and jitter ratio σRJ /ADJ (b) of fitted test
distributions over varying shape σRJ /ADJ , At ≡1, N =108 , R=Rsim =3.3·105 .
possible value of N to obtain a pessimistic amplitude. That is, fitted tail amplitudes become
smaller with increasing N , which is also an effect of the asymptotic tail behavior in Q-domain.
ln(σ t /σRJ )
(figure 5.4(a)) (5.5b)
y= σt /(2µt )
ln (figure 5.4(b)) (5.5c)
σRJ /ADJ
where the inverse can be determined iteratively using a simple Newton iteration. Equivalent to
tail amplitudes, regression coefficients are given in table 5.2, for three important sample size can-
didates N ={105 , . . . , 109 }. Regressions are carried out in the interval σRJ /ADJ =[2·10−3 , 0.5],
where equation (5.5c) additionally uses y≤ln(101 ) as upper bound. This is to guarantee for suf-
ficient accuracy of fitted polynomials (r-squared statistic 1−r2 <10−3 , defined in section 5.4) and
does not restrict the analysis, since one is only interested in finding minimum parameter values.
Note, that instead of the empirical equations (5.3) and (5.5) it is always possible to rely on fast
numerical approximations as described initially. They allow to quickly investigate the average
fitting behavior for arbitrary N and R. In the subsequent sections the numerical approximations
will be used together with additional equations to guarantee for correct behavior of fitting methods.
A corresponding design example is also given in section 5.5.
63
5. H ARDWARE D ESIGN A SPECTS
TABLE 5.2.: Coefficients for equations (5.5b) and (5.5c), valid for QN and sQN methods.
64
5.2. M INIMUM S AMPLE S IZE
included for regression analysis, and must at least range over two decades. Results also showed,
that ∆Pt should be selected as large as possible in order to include the complete visible RJ tail.
This second problem is also depicted in figure 5.5 for a right bathtub curve with minimum tail
amplitude At,min . Since only part of the Gaussian RJ is observed at the TJ distribution ending,
At,min must significantly exceed the upper probability level Pup imposed by ∆Pt . Therefore, as a
rule of thumb equation (4.20) provides the following constraints:
where Pup is the product of multiplication factor ∆Pt and probability granularity 1/N . The latter
condition is rather vague and does not indicate the minimum amplitude as precise as desired. Since
∆Pt is a free design parameter, we like to include all of the observable RJ tail for the worst case
situation in figure 5.5. However, the shape of a measured distribution is basically unknown and
thus, there is no information available on how deep Pup must be located below At,min .
In a simple conservative approach a pessimistic
threshold can be selected to define the observable CDF
tail part of a Gaussian function. One can for ex- At,min
RJ tail
ample assume that the Gaussian function is at least
visible up to the lower inflection point, which is Pup
located at one standard deviation from the mean ∆Pt
value. In order to determine the tail part, thus 1/N
the tail probability p at the inflection point must
be calculated. Therefore we recall the Gaussian
probability function or CDF from equation (3.9) F IGURE 5.5.: Right bathtub curve (solid)
with fitted Gaussian tail
−1 1 −q (dashed).
CDF(q) = Q (q) = p = · erfc √ (5.7)
2 2
which is also the inverse of the Q-function (3.10). The inflection point of a Gaussian is located at
x=µ − σ. In Q-domain, the Gaussian is a linear function with the standardized variable q:
x − µ x=µ−σ
q= −−−−→ q = −1 (5.8)
σ
The corresponding tail probability is thus:
It defines the probability level where the observable Gaussian part of a distribution tail starts. If
referred to the minimum amplitude At,min , one can specify a conservative threshold for Pup :
This final result can be used together with equation (5.3), which allows to identify At for a given
distribution shape. Thus, the distribution shape can now also be related with a minimum sample
size requirement.
Note that the obtained condition is only valid if fitted tails truly follow a Gaussian below the
inflection point. According to the central limit theorem, the combination of Gaussian RJ with a
65
5. H ARDWARE D ESIGN A SPECTS
bounded DJ component (section 3.2.2) always leads to a TJ distribution which is more Gaussian-
like. Therefore, the synthesis principle already indicates validity, and in fact, empirical analyses
with the scaled Q-normalization method confirm this assumption. As long as the algorithm is able
to correctly identify the tail part, the fitted region also includes the inflection point.
However, this cannot be generalized for arbitrary distribution shapes and thus, the conservative
threshold may have to be re-specified with more pessimistic tail assumptions. Unfortunately,
no matter how pσ is selected, a certain risk to choose it inappropriately must always be faced.
Although rather theoretical, this is still an inherent disadvantage of the ĉ1.2 fitting algorithm from
section 4.3.
In section 4.4 the constant threshold scenario Qth,c was introduced as an alternative to define
the Gaussian tail part in Q-domain (figure 4.27). The worst case analysis situation is here given
by the shortest Q-tail, where the distance between lower tail end and the threshold Qmin becomes
minimum. In this case, the fitted tail suffers from the highest amount of statistical random varia-
tion.
In mathematical terms one can say that Qmin must be located above the required minimum
interval for tail fitting, which is given by ∆Pt /N when transformed into scaled Q-domain. The
worst case situation is given with a maximum scaling factor kmax , and thus
∆Pt
Q kmax · ≤ Qmin (5.11)
N
where the left hand side describes the ∆Pt /N interval in scaled Q-domain. Thus, by applying
Q−1 on both sides and by inserting the minimum amplitude At,min =1/kmax we have
∆Pt
≤ Q−1 (Qmin ) (5.12)
N · At,min
which can be rewritten as
∆Pt
At,min ≥ (5.13)
Q−1 (Q min ) ·N
This final result is very useful as it directly relates the sample size N with the minimum tail
amplitude At,min . Since N describes the measurement effort, the minimum amplitude At,min
returns the resulting benefit. The two parameters ∆Pt and Qmin influence the relation and are
now both included with the Qmin based optimization scenario. That means, equation (5.13) was
originally derived for the optimization scenario with Q-domain threshold Qmin from section 4.4.
However, now it also shows the missing link for the conservative probability threshold ∆Pt in
equation (5.10).
This threshold was needed to derive an amplitude relation for the ĉ1.2 optimization scenario in
section 4.3, using only the fitting parameter ∆Pt . Now with the same structure of the formula we
can easily see, that the probability factor pσ corresponds to
This means, the additional Qmin parameter transforms the conservative approach of equation (5.10)
into the exact relation (5.13). We recognize this as a beneficial property of the Qth,c algorithmic
scenario from section 4.4.
In order to optimize equation (5.13) with respect to minimum tail amplitude, we choose ∆Pt as
small as possible, using the minimum ∆Pt =102 (equation (5.6)). The threshold parameter Qmin
is back-transformed into probability domain, where its reverse value also influences At,min . In
figure 5.6 the Q−1 behavior is displayed as a function of Qmin . According to this plot one would
like to minimize Qmin , but this comes along with a degraded performance since an optimum value
was already identified at Qmin =−1.2 (see figure 4.31). Hence, as long as At,min is not a critical
66
5.3. M INIMUM T IME R ESOLUTION
Q−1(Qmin)
As an example, in section 4.4.3 distributions
were analyzed using N =107 and Qmin =−1.2 −2
10
as threshold. Now we also like to guarantee a
minimum tail interval to avoid outliers caused
by statistical tail variations. With the minimum −3
10
∆Pt =102 and equation (5.13): 0 −0.5 −1
Qmin
−1.5 −2 −2.5
The result allows to correctly fit tail amplitudes down to At,min =8.7·10−5 , without causing large
outliers. For the worst case with quadratic curve shaped DJ and equation (5.4) we yield a smallest
jitter ratio of σRJ /ADJ =6.0·10−3 , which covers nearly all of the test distributions in figures 4.26
and 4.33. Thus, none of the investigated distribution shapes is strongly affected by outliers. How-
ever, ∆Pt should always be selected as large as possible to guarantee for an optimum outlier sup-
pression. Further, the obtained result is only valid for simulations where R=Rsim and thus, when
time resolution does not affect the performance. This problem domain is addressed subsequently.
67
5. H ARDWARE D ESIGN A SPECTS
regression analysis. Thus, again the two algorithmic approaches ĉ1.2 and Qth,c with ∆Pt interval
and Qmin threshold must be considered separately.
68
5.3. M INIMUM T IME R ESOLUTION
Emed
Emed
E
0 −2 −1 0
0 −2 −1 0
0 −2 −1 0
10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
Emed
Emed
E
0 −2 −1 0
0 −2 −1 0
0 −2 −1 0
10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
F IGURE 5.8.: Smallest analyzable RJ component σt,min . Empirical relation (5.20) is veri-
fied with respect to varying time resolution R={1024, . . . , 32} and sample size
N ={105 , 106 , 107 }. ∆Pt ={102 , 103 , 104 }, K=250, uniform DJ type, ĉ1.2 opti-
mization scenario.
∆ Pt / N = 10−6
∆ Pt = 101
3
σt,min × R
0.5
∆ Pt = 106
∆ Pt / N = 10−1
0.3 4 5 6 7 8 9
10 10 10 10 10 10
N
F IGURE 5.9.: ∆Pt selection chart for identifying σt,min ·R, as described by equation (5.20).
∆Pt ={101 , . . . , 106 } (dashed), ∆Pt /N ={10−6 , . . . , 10−1 } (solid).
69
5. H ARDWARE D ESIGN A SPECTS
In figure 5.9 additionally, a chart for (5.20) is provided to help selecting the free design param-
eter ∆Pt , and to verify whether a certain minimum RJ tail can be fitted correctly. For example,
knowing N and R one can easily verify whether a desired σt,min is guaranteed for a certain ∆Pt .
σt,min decreases linearly with the number of bins, and is thus represented by the normalized,
dimensionless variable σt,min ·R. The chart is constructed using two different types of curves,
where either ∆Pt or the ratio ∆Pt /N is constant. This is to guarantee the two requirements with
respect to outlier suppression (equation (5.6)) and minimum tail amplitude (equation (5.10)) in-
dependently. Both curve types allow for the analysis of a smaller RJ standard deviation, if ∆Pt
is increased while N is constant. The maximum ∆Pt is only restricted by the minimum tail am-
plitude. Note that in equation (5.20) the value ∆Pt /N forms the upper probability level for tail
selection. If constant, the sample size N can be used to increase the fitting region and thus, to
identify a smaller RJ standard deviation.
An interesting effect is further noticed with constant ∆Pt . If more jitter samples are used for
collecting distributions, σt,min becomes larger. This seems contradictory, but is a result of the
nonlinear Q-function behavior when only the lowest probability region is used for tail fitting. In
fact, the benefit in this case lies in a smaller minimum amplitude At,min . To summarize these
observations, both ∆Pt and N should be chosen as large as possible, without violating equa-
tion (5.10).
The upper bound Qmin must belong to the linearized Q-tail part, which can easily be verified by
equation (5.13). Again, it is not possible to specify a tail amplitude At because it depends on the
distribution shape, but the minimum amplitude At,min from equation (5.13) can be used instead.
This is because correct convergence for the fitting method is assumed, and At,min forms a worst
case scenario which allows for pessimistic σt,min estimation. Thus, when inserting equation (5.13)
into (5.22) :
2 1
σt,min ≤ · −1 (5.23)
R Q
Q (Qmin )
−Q
∆Pt min
where the σt,min function is now reduced to a simplified form without depending on the sample
size N . This expression cannot be reduced further, due to the non-linearity of the Q-function.
Equivalent to the previous analysis, applicability of this equation is demonstrated in figure 5.10.
Vertical solid lines mark the calculated values from equation (5.23). The course of median er-
ror Emed over varying jitter ratio shows where the fitting algorithm starts to fail at each of the
three sample sizes N ={105 , 106 , 107 }. The Q-domain threshold Qmin =−1.0 is here used for
70
5.3. M INIMUM T IME R ESOLUTION
Emed
Emed
0.02 0.02 0.02
0 0 0
−0.01 −2 −1 0
−0.01 −2 −1 0
−0.01 −2 −1 0
10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
Emed
Emed
0.02 0.02 0.02
0 0 0
−0.01 −2 −1 0
−0.01 −2 −1 0
−0.01 −2 −1 0
10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
F IGURE 5.10.: Smallest analyzable RJ component σt,min . Empirical relation (5.22) is veri-
fied with respect to varying time resolution R={1024, . . . , 32} and sample size
N ={105 , 106 , 107 }. Qmin =−1.0, ∆Pt =102 , K=250, uniform DJ type, Qth,c
optimization scenario.
1
Qmin = −2
σt,min × R
Qmin = 0
0.5
0.3 1 2 3 4 5 6
10 10 10 10 10 10
∆ Pt
F IGURE 5.11.: ∆Pt and Qmin selection chart for identifying the normalized variable σRJ,min ·R,
as described by equation (5.23). Qmin ={0, −0.5, −1.0, −1.5, −2.0}.
71
5. H ARDWARE D ESIGN A SPECTS
calculations, which is different from the optimum value in section 4.4.2. This is due to the coarse
time resolutions RRsim , and allows to include more tail samples with the Qth,c optimization
scenario.
Again, σt,min values in equation (5.23) refer to the fitted tail parameter of a TJ distribution, and
can only be plotted in figure 5.10 when using the inverse of equation (5.5b). This yields different
results for the three sample sizes, which are located very close to each other. Here only the N =107
case is shown, due to more pessimistic values.
If ∆Pt is increased, σt,min estimates become smaller, which allows to search for a suited param-
eter value without depending on the sample size N . The normalized standard deviation σt,min ·R
is plotted in a chart for ∆Pt selection over different Qmin values (figure 5.11). The chart suggests
∆Pt to be selected as large as possible, but note that equation (5.13) must always be fulfilled.
Also, ∆Pt ≥102 is highly recommended to guarantee for sufficient outlier suppression.
In order to derive empirical relations, first the complexity of this four-dimensional function must
be reduced. As an introductory example, in figure 5.12 the estimation loss EL over varying dis-
tribution shape is depicted, by combining RJ (σRJ ) with uniform type DJ (ADJ ). TJpp estimates
are obtained from fitted bathtub tails using different time resolution R. In the example, the sQN
algorithm uses the optimized parameter configuration from section 4.3.2, with ĉ1.2 and ∆Pt =105 .
Basically two effects are noticed. First, the algorithm is highly biased if random jitter falls
below a certain minimum σRJ,min . This limiting effect is caused by the discrete time resolution
and has already been dealt in the previous section. Second, EL increases when either σRJ or R are
reduced at constant jitter ratio σRJ /ADJ . That means, the surfaces in figure 5.12 are symmetric
and thus, the ratio σRJ /ADJ can be reused as simplifying shape variable. However, an additional
dependency on σRJ and R is observed, combined with a high error ripple if either one of the two
variables is varied.
In order to quantify the estimation error, equation (5.24) must be simplified. Considering again
the jitter ratio σRJ /ADJ as single variable, only the shapes of largest error from the analysis in
figure 4.26 may be investigated within the scope of a worst case analysis. This yields a single shape
value per DJ type. Additionally, with constant ratio both σRJ and ADJ vary simultaneously, so
that each parameter also represents the total distribution size. The estimation error thus reduces to
72
5.4. E STIMATION E RROR A NALYSIS
EL
EL
0.01 0.01 0.01
0 0 0
−3 −3 −3
10 10 10
2 2 2
10 10 10
−2 0 −2 0 −2 0
10 10 10 10 10 10
−2 −2 −2
10 10 10
−1 −4 −1 −4 −1 −4
σRJ 10 10 σ /A σRJ 10 10 σ /A σRJ 10 10 σ /A
RJ DJ RJ DJ RJ DJ
5
(a) R=Rsim =3.33 · 10 (b) R=1024 (c) R=512
0.04 0.04
0.03 0.03
0.02 0.02
EL
EL
0.01 0.01
0 0
−3 −3
10 10
2 2
10 10
−2 0 −2 0
10 10 10 10
−2 −2
10 10
−1 −4 −1 −4
σRJ 10 10 σRJ/ADJ σRJ 10 10 σRJ/ADJ
F IGURE 5.12.: Estimation loss EL for different values of σRJ , ADJ , and R. Rsim in fig-
ure 5.12(a) is obtained with a simulator time resolution of 1 fs in a 3Gb/s in-
terface. N =108 , ∆Pt =105 , K=250, ĉ1.2 optimization criterion.
TABLE 5.3.: Selected worst case shape values σRJ,min /ADJ and corresponding worst case tail
amplitudes At,min (N =108 ) for different DJ types.
The variable σRJ now describes the overall distribution size, while ADJ is discarded. Remember,
that ADJ does not fully disappear as parameter, it has only been transformed into a dependent
variable. Varying the parameter σRJ now also means to vary ADJ according to the selected worst
case ratio.
This jitter ratio depends on the selected DJ type and thus, has to be determined for each of
the investigated shapes. From the performance results in sections 4.3.3 and 4.4.3 we can easily
determine these ratios for subsequent analysis. They are given in table 5.3 together with worst
case tail amplitudes, obtained by equation (5.3) at N =108 . The shape values have been selected
as power-of-two to simplify the performance analysis, so that the numerical approximations of
true total jitter values TJpp,true can be reused.
The next step is to search for a linear or logarithmic dependency between two of the independent
73
5. H ARDWARE D ESIGN A SPECTS
IQR
EL
0.01 0.01 0.02
0 0 0
−3 −3 −3
10 10 10
1 1 1
10 10 10
−2 −2 −2
10 2
10 10 2
10 10 10
2
3 3 3
σ
−1
10 10 σRJ,min
−1
10 10 σ
−1
10 10
RJ,min R R RJ,min R
IQR
EL
0.015 0.015 0.03
0 2 0 2 0 2 1 0
1 0 1 0
10 10 10 10 10 10 10 10 10
σRJ,min ⋅ R σRJ,min ⋅ R σRJ,min ⋅ R
F IGURE 5.13.: Median estimation error Emed (a,d), interquartile range IQR (b,e) and estima-
tion loss EL (c,f) over varying R and σRJ,min . N =108 , ∆Pt =105 , K=250,
σRJ,min /ADJ =const.=1/4, uniform DJ type.
variables from equation (5.25), while keeping the third one constant. As depicted in figure 5.13,
such a relation is found for a constant sample size N . If the parameters R or σRJ are varied, all
three performance indicators Emed , IQR and EL yield plane surfaces that consist of a convergence
region, which is to some degree affected by ripple. Inside the convergence region, bathtub tails
are supported by sufficient bins, thus allowing the fitting algorithm to correctly extrapolate tails.
When either σRJ or R become too small, the algorithm shows a large error bias. The ripple
increases when moving toward the convergence limit, which is an effect caused by the coarse time
discretization and will also be investigated later on in this section.
The regression analysis can be simplified by changing the variable representation into the prod-
uct form σRJ ·R, as demonstrated in the bottom row of figure 5.13. This is due to the constant
ratio σRJ /ADJ , which changes the whole distribution size when only σRJ is varied. The fitting
algorithm cannot distinguish between a variation of the distribution size or the time resolution. In
fact, the extrapolation error is only influenced by the number of bins which form the bathtub tail.
An increase in performance of the algorithm can therefore be achieved with a larger number of
bins R as well as a larger distribution size given in terms of σRJ .
As a result, another dependency between the variables R and σRJ has been identified. One
of them can be discarded if only the product σRJ,min ·R is considered, which further reduces the
estimated error in equation (5.25) to a function of two variables:
max{Emed , IQR, EL } = f (σRJ ·R, N ) , σRJ,min /ADJ = const. (5.26)
Consequently, for a constant sample size N , estimation performance in figures 5.13 (d,e,f) can also
be approximated with a simple regression line.
The variable product σRJ ·R is here referred to as node or bin density of a distribution. When
comparing this density with parameter selection charts from figures 5.9 and 5.11, we notice that
it also corresponds to the normalized standard deviation. Depending on the selected algorithmic
type, equations (5.20) and (5.23) can thus be used to identify the expected convergence limit.
74
5.4. E STIMATION E RROR A NALYSIS
−1 −1
10 10
−2
Emed
IQR
−2
10 10
−3 −3
10 10
0 0
10 5
10 5
10 10
6 6
1 10 1 10
10 7 10 7
10 10
8 8
σRJ,min × R 10 N σRJ,min × R 10 N
(a) (b)
F IGURE 5.14.: Surfaces for empirical analysis of Emed (a) and IQR (b). The worst case
shape σRJ /ADJ =1/4 (uniform type DJ) is here investigated in the range
N ={105 , . . . , 108 } and σRJ ·R={0.8, . . . , 51.2}, K=250.
In figure 5.13 the ĉ1.2 based algorithm was used with N =108 and ∆Pt =105 as parameters. The
selection chart in figure 5.9 and equation (5.20) yield σt ·R=0.79 as limit. With σRJ /ADJ =1/4
and equation (5.5b) we get σRJ ·R=0.72, which is consistent with the observed surfaces in fig-
ure 5.13.
Using the sample size N as second independent variable, the estimation error can be represented
by two-dimensional surfaces. If regression planes achieve acceptable accuracy, an empirical de-
scription of the extrapolation error thus becomes possible. Subsequently, the two algorithmic
candidates from sections 4.3 and 4.4 are investigated separately.
75
5. H ARDWARE D ESIGN A SPECTS
TABLE 5.4.: Emed , IQR and EL regression coefficients for equation (5.30). ∆Pt /N =10−3
(planes for large N ) and ∆Pt /N =10−2 (planes for small N ), ĉ1.2 algorithm, pa-
rameter intervals specified above, K=250.
76
5.4. E STIMATION E RROR A NALYSIS
TABLE 5.5.: Emed , IQR and EL coefficients for equation (5.30). Qmin =−1.0, ∆Pt =102 , Qth,c
algorithm, parameter intervals specified above, K=250.
of these two variables. At very low densities, ∆Pt =102 guarantees for sufficient bins to be in-
cluded with regression analysis. This avoids outliers, but unfortunately the pessimistic estimation
property with positive error bias for Emed is not guaranteed anymore. In fact, Emed now also
returns negative values, as can also be seen in figure 5.10. This is an effect caused by statistical
tail variations, which leads to highly overestimated scaling factors and optimistic TJ values.
The same original grids are used for the regression planes as with the previous analysis. Ta-
ble 5.5 contains the regression coefficients a0 , a1 and a2 for each of the investigated DJ types,
as well as the r-squared statistic to indicate the quality of fitted planes. The empirical relation is
again given by equation (5.30). This time, Emed planes highlight significantly smaller r2 values,
because of the partial influence of negative errors at lowest bin densities. This problem is over-
come by selecting a smaller parameter range of σRJ ·R for planes with small N . However, besides
the pure Gaussian case, EL is significantly larger compared to the ĉ1.2 based algorithm.
To highlight this difference in a brief example, we assume a jitter measurement system with the
parameters R=128 and N =108 . For the sQN method with ĉ1.2 based tail fitting ∆Pt =105 is se-
lected, and a minimum analyzable standard deviation σt,min =6.2·10−3 obtained (equation (5.20)).
According to equation (5.5b), at a worst case jitter ratio of σRJ /ADJ =1/4 (uniform DJ) we get
σRJ,min =5.7·10−3 . The worst case error is determined using equation (5.30) and yields
Emed = 1.46%, IQR = 1.64%, EL = 3.90% (5.31)
For the Qth,c based tail fitting with Qmin =−1.0 instead
Emed = 1.31%, IQR = 2.42%, EL = 4.98% (5.32)
This result shows that the Qth,c algorithm, although less biased, suffers from a significantly larger
statistical spread which is especially noticed with EL . A similar error characteristic is observed for
all four DJ types. Also for the pure RJ case, the error of the Qth,c algorithm becomes excessively
large as soon as σRJ ·R≤5 and N ≤106 . Generally, this makes the ĉ1.2 algorithm a better suited
choice for tail fitting with hardware based jitter measurements. Therefore, only the ĉ1.2 algorithm
will be further utilized subsequently.
77
5. H ARDWARE D ESIGN A SPECTS
4
10
−1
10
5
10
Emed
−2
10
6
10
N
−3
10
4 7
10 10
6
10
N
8
10
8 1 1.6
10 0.5 0.5 1 1.6
σRJ,min × R σRJ,min × R
(a) (b)
F IGURE 5.15.: Error ripple effect: simulated Emed surface (a) and expected “error valleys”
according to equation (5.35) (b) with a test distribution σRJ,min /ADJ =1/4
(uniform DJ). The investigated parameter ranges are N ={104 , . . . , 108 } and
σRJ,min ·R={0.4, . . . , 1.6}. K=250.
78
5.4. E STIMATION E RROR A NALYSIS
in order to guarantee that the fitting algorithm operates in a minimum error region.
As a certain drawback, the Gaussian model parameters (At , µt , σt ) must be known. In a simu-
lation they can be identified using a numerical approximation of the ideal bathtub function, equiv-
alent to the calculation of tail parameters in section 5.1. At N =106 thus, we yield At =0.380,
µt =0.0653 and σt =0.0531. Note, that these values form a compromise, since {At ,µt ,σt }=f (N ),
due to the asymptotic tail behavior in Q-domain.
In a real measurement scenario Gaussian model parameters can be approximated using the
median values of multiple tail fits. The remainder function (5.36) is then an indicator on how far
the tail edge is located from the outermost distribution sample, and hence, tells whether the fitting
result lies within an error maximum or minimum. Equation (5.35) can additionally be inverted to
identify suited values for the sample size N .
1
N= (5.37)
2n/R−µt
At · Q−1 σt
where n defines the trace number to be located on. If N is adjustable, one is thus able to move
from an error maximum to an error minimum by simply changing the sample size. However, this
analysis is only valid for a single distribution shape, and cannot be applied to a broad range of
distributions. Further the measurement system must not be affected by differential non-linearity
(DNL) error, which is rarely the case. DNL is caused by timing mismatches or process variations,
and leads to a smoothing of the rippled surface from figure 5.15 as well as a large statistical spread
of fitting results. This effect is investigated subsequently.
N (1/R, σDN L )
0 1/R 2/R
−2/R −1/R
F IGURE 5.16.: DNL error model to describe the effect of process variations.
of an ideal measurement system. As an additional problem, the DNL error is summed up over the
delay line and yields an integral non-linearity (INL) which significantly exceeds the DNL values.
This effect can only be reduced if the PLL output clock is very clean and directly provides multiple
phases.
Similar to the empirical error analysis, the DNL error term is included as third variable and
yields a hyperplane where multiple linear regression can be performed. In order to deal with the
large computational demand, a reduced grid resolution is chosen with 25×25 nodes for bin density
σRJ ·R={0.8, . . . , 51.2} UI, as well as N ={5·105 , . . . , 108 } and N ={104 , . . . , 106 } respectively.
Each plane is simulated with respect to varying DNL error in the range σDN L ={0.0, . . . , 0.19}
using an equally spaced distance of 0.01.
79
5. H ARDWARE D ESIGN A SPECTS
TABLE 5.6.: Coefficients for Emed , IQR and EL with included DNL error, equation (5.38).
∆Pt /N =10−3 (large N ) and 10−2 (small N ), ĉ1.2 algorithm, K=200.
For multiple linear regression a suitable model description was identified with the relation
y = −a0 − a1 x1 − a2 x2 − a3 x3 − a4 x2 x3 (5.38a)
x1 = ln(N ), x2 = ln(σRJ · R), x3 = ln(1 + σDN L ) (5.38b)
y
{Emed , IQR, EL } = e (5.38c)
where the last coefficient a4 also considers a correlation between bin density and DNL error, which
proved to significantly increase the quality of fitted hyperplanes. The coefficients are given in
table 5.6. Here, smaller r-squared values for the fitted hyperplanes of Emed are especially noticed
with the pure RJ shape. This is, because the DNL error highly affects statistical tail variations
and pure Gaussian distributions suffer from the strongest influence of random noise. Thus, fitted
hyperplanes also become inaccurate.
80
5.5. D ESIGN E XAMPLES
81
5. H ARDWARE D ESIGN A SPECTS
For the selected minimum variance σRJ,min , the resulting coarse bin density is the major contrib-
utor to the overall error. This influence becomes even more evident if additionally a DNL error of
σDN L =0.05 UI is assumed and the empirical equation (5.38) with table 5.6 applied:
which demonstrates that the statistical spread of the fitting method can be highly affected by DNL
error. However, this is also an effect of low bin densities, since DNL has been modeled as a random
step with time interval 1/R (section 5.4.3), and can thus also be reduced using a higher time
resolution. This design example will also be continued in section 6.5 to compare the extrapolation
error of the sQN method with the conventional Q-normalization (QN) method.
(N · R) · R N · R2
tt = = (5.45)
C · fs C · fs
The sQN fitting method shall operate appropriately without being affected by large error oscilla-
tions and thus, σRJ ·R≥2 is required (also see figure 5.15). With a selected number of bins R=91,
this can be guaranteed if the minimum standard deviation σt,min ≥σRJ of the fitted random jitter
component is greater or equal to
Note that the system is still able to operate down to the minimum value given by equation (5.20),
but additionally suffers from error oscillations if σRJ ·R<2.
As shown in equation (5.45), the test time tt can be linearly decreased with a larger number of
parallel counters C. Equivalent, the number of samples N can also be increased to improve accu-
racy of the jitter analysis method, if tt is kept constant. Thus, the expected worst case extrapolation
error EL of the sQN method can be plotted against the number of implemented parallel counters
C, which is a direct measure for the hardware expense. Therefore, empirical relation (5.30) and
the table of coefficients 5.4 are applied, together with the regression planes for small sample sizes.
In the calculations, we assume the pure Gaussian RJ case as well as sinusoidal DJ, where the latter
is typically observed with high-speed PLLs that are affected by spectral spurs.
As a result, the extrapolation error EL in figure 5.17 is depicted over increasing number of
counters C, which also defines the sample size N . N is determined from the inverse of equa-
tion (5.45), while a minimum bin density of σRJ ·R=2 is assumed. For the pure RJ case with
C≥14 and tt =20ms (⇒N ≈26k), the given measurement system is for example able to estimate
the TJ of the PLL under test with <15% error. Note that this result reflects the combined influence
of worst case error bias and spread, and includes approximately 97.5% of estimates. The presented
design example will also be continued in section 6.5 for comparison with the QN method.
82
5.6. S UMMARY
0.5
0.4
0.3
tt=20ms
Estimation Loss
Sinusoidal DJ
0.2
tt=20ms tt=50ms
0.1
Only RJ
tt=50ms
1 10 100
Number of Counters
F IGURE 5.17.: Estimation loss EL of sQN method over varying number of counters C, with
tt ={20, 50} ms, and worst case sinusoidal DJ (red) or pure RJ (black) case.
5.6. Summary
Hardware related design aspects were investigated to utilize the scaled Q-normalization method
for on-chip jitter diagnosis or together with built-in jitter measurement (BIJM) system. Influences
of limited sample size N as well as number of bins R on the algorithmic performance were inves-
tigated. For each analysis, the two algorithmic scenarios ĉ1.2 (section 4.3) and Qth,c (section 4.4)
were investigated independently.
In order to characterize the tail parameters of fitted test distributions, first, the polynomial equa-
tions (5.3) and (5.5) were derived. These allow to change between the variable representation
prior (σRJ , ADJ ) and after (At , σt , µt ) distribution synthesis. The coefficients in table 5.1 are
used together with equation (5.3) and specify the tail amplitude At obtained with the sQN fitting
method. The coefficients in table 5.2 and equation (5.5) specify minimum requirements of the
sQN method with respect to the tail parameters σt and µt . The obtained results are also valid for
the conventional Q-normalization (QN) method without scaling (chapter 6) and thus, allow for
parameter specification of both methods.
With the ĉ1.2 algorithm, the minimum tail amplitude At,min is estimated by defining a conser-
vative threshold, as shown with equation (5.10). The Qth,c fitting algorithm instead allows for
the derivation of an exact equation (5.13). This result also highlights the missing link to the first
algorithm.
The time resolution variable divides the unit interval into a discrete number of bins R, and
causes a limiting effect for maximum tail slope, which can be expressed as minimum Gaussian
standard deviation σt,min . With ĉ1.2 , equation (5.20) has been derived to identify the σt,min value
which can be fitted correctly by the algorithm. Validity of this equation has been demonstrated em-
pirically, and a selection chart for ∆Pt has been given in figure 5.9, in order to simplify a suitable
choice. The two conditions in equations (5.6) and (5.10) further guarantee for sufficient outlier
suppression as well as a robust algorithmic behavior. With the Qth,c algorithm, equation (5.23)
has been derived to give a pessimistic estimate for σt,min . A selection chart has been provided in
figure 5.11 as well. As a clear advantage of this second algorithm, Qmin and ∆Pt can be adjusted
independently from the sample size N . Correct convergence of the algorithm is again guaranteed
with ∆Pt as large as possible and condition (5.13) fulfilled.
Considering the combined influence of sample size and time resolution on the extrapolation
error of fitting algorithms, empirical relations were derived to approximate error bias and spread
as a function of sample size N and bin density σRJ,min ·R (see equation (5.30)). These empirical
relations investigate the worst case distribution shapes of the four important DJ types defined in
section 3.2.2, as well as the pure Gaussian RJ case. They are meant to aid the designer in finding
an optimum performance trade-off. The corresponding empirical coefficients can be found in
83
5. H ARDWARE D ESIGN A SPECTS
table 5.4 for ĉ1.2 , and in table 5.5 for the Qth,c algorithm. According to obtained results, the
ĉ1.2 algorithm from section 4.3 clearly highlights a better performance. This is mainly due to the
unfavorable behavior of Qth,c at lowest bin densities, which leads to large error oscillations and
may even cause negative errors. This highly degrades the extrapolation performance, and also the
quality of fitted regression planes.
The observed error oscillations, or error ripple effect at very small bin densities σRJ ·R<2 has
been investigated as well in section 5.4.2. As a fundamental result, equation (5.35) describes the
observable oscillations, and can be used to identify suitable working regions in order to avoid error
maximums. However, results behave optimal only for a single distribution shape, which is rather
impractical.
In the case of differential non-linearity (DNL) error, as caused by timing mismatches of the jitter
measurement system, the statistical spread of TJpp estimates increases significantly and becomes
the major contributor to overall estimation loss. Thus a DNL error model together with empirical
equation (5.38) and the table of coefficients 5.6 has been derived. It describes the influence of
DNL error on estimation performance in terms of the standard deviation σDN L .
In a first design example, applicability of the derived equations with respect to jitter diagnosis
has been demonstrated. Starting with a worst case RJ of σt,min =0.01UI and a time discretiza-
tion R=128, the minimum amplitude At,min was determined by assuming ISI dominated jitter
(uniform DJ). This allowed to correctly specify the design parameter ∆Pt , in order to guarantee
for correct convergence of the algorithm. With equation (5.30) and the table of coefficients 5.4
a worst case error bias Emed ≤2.0% as well as an estimation loss EL ≤5.4% were guaranteed if
the measurement system was not affected by DNL. Otherwise, with an assumed DNL standard
deviation of σDN L =0.05UI and equation (5.38) together with the table of coefficients 5.6, these
values increased to Emed ≤2.3% and EL ≤7.4% respectively.
Finally in a second example, the derived empirical relations were applied to a jitter measurement
system for fast production testing of high-speed PLLs. The sQN method was able to estimate the
true TJ budget in a given measurement time of tt =20ms with less than 14.9% error (pure RJ case)
as well as 13.5% error (sinusoidal DJ), if only the number of counters was increased to C=14 in
order to achieve a larger sample size N .
Parts of this chapter have also been published in [C4,C8].
84
6. Comparison of Gaussian Tail Fitting
Methods Based on Q-Normalization
In this chapter the performance of the scaled Q-normalization (sQN) method is compared with
various other tail fitting principles based on the Gaussian quantile normalization. This analysis
is meant to give useful insight to the performance and stability of fitting algorithms, and to high-
light their advantages and drawbacks. So far, literature is still missing on such comparisons and
generally lacking from a detailed performance description. This is partly also due to the high
computational demand associated with statistical evaluations. In this work, this problem is dealt
by a powerful cluster of up to fifty parallel workstations using 3GHz Intel Xeon processors.
From the variety of histogram based fitting techniques [51,54,84,95,124,136] this chapter only
focuses on methods related to the Gaussian Q-normalization principle. Obviously, a chi-squared
test as for example used in [52,84,90] would be a prominent candidate, but is omitted here because
of the highly individual structure of such algorithms. In fact the optimization process is quite
complex and typically includes histogram smoothing, outlier removal, tail part identification and
a Gaussian model search over several optimization stages. Implementing a chi-squared test with
acceptable accuracy and robust behavior is thus time consuming and hard to achieve. In contrast,
the presented fitting methods do not require any data preprocessing, and can directly be applied
onto raw jitter distributions.
The first type of fitting algorithm being compared against the sQN method is the conventional
Q-normalization (QN) method without scaling. The algorithm is simply obtained by omitting the
pre-scaling factor k and directly performs a linear regression analysis in Q-domain, as already
mentioned in section 3.1.2. This principle was proposed in [51, 111] and subsequently also de-
scribed in [82, 123]. The second class of algorithms uses higher order polynomials for tail fitting
in Q-domain, and was first suggested by Hong in [54]. The idea is to replace the linear regression
stage by a higher order polynomial regression, which fits polynomial functions into the measured
Q-tails. The obtained polynomial coefficients thus describe a parameterized bathtub function,
which can easily be used to extrapolate distributions down to the BER level of interest and thus,
to recover the TJ timing budget. This principle allows for a whole class of polynomial regression
methods (QP2, QP3, . . .) to be compared against the sQN method.
In figure 6.1 the optimization scheme for Gaussian Q-normalization, when combined with poly-
nomial regression, is depicted. Polynomials of arbitrary order can be fitted to the distribution tails
in Q-domain, where the resulting regression coefficients {a0 , . . . , ap } denote the functional rela-
tion between jitter amplitude x and quantile q. Similar to the sQN method, the regression error
σ̂err (n) can again be interpreted as a function of tail length n, when starting the regression with
CDF(x) error
Polynomial Reg.(n)
Q(· · · ) order p
σ̂err (n)
q = f (a0 + a1 x + ... + ap xp )
F IGURE 6.1.: Optimization scheme for Q-normalization combined with polynomial regression.
85
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
outermost tail samples and recursively moving toward higher probabilities. Therefore, the basic
implementation principle of the fitting algorithm remains the same.
Note, that for the special case of order p=1, the regression function reduces to a line with
offset o=a0 and slope s=a1 . Since the scaling factor k is not included, this corresponds to the
conventional method without scaling (QN), as described before. In other words, the QN method
is equal to the first order polynomial regression (QP1) in Q-domain.
When trying to combine the scaling factor k with higher order polynomials, the optimization
becomes unstable and diverges toward meaningless values of k<1. Thus, only for the linear case
with p=1, the scaling factor k can be part of the optimization scheme. This also confirms the
sQN method as three dimensional approach to Gaussian tail parameter search. The QN method
instead, with k=A=1 always fits a Gaussian function of maximum amplitude, and can thus only
be used to retrieve mean µ and standard deviation σ of the Gaussian model. Further, higher order
polynomials (QP2, QP3, . . .) cannot be used at all for retrieving Gaussian model parameters.
In the subsequent sections first, an efficient algorithmic implementation as required for the
polynomial regression of Q-tails, equivalent to the realization of sQN in section 4.1.4 is described.
Then optimum selection criteria for conservative fitting parameters Pt,min and ∆Pt are again dis-
cussed. This is meant to improve the robustness of polynomial fitting methods by selecting an
appropriate tail region. Starting with a performance evaluation of the different methods, poly-
nomials up to the fourth order are investigated, which is sufficient as will also be shown. Then
the polynomial methods are compared with sQN, and as a result it is shown that the conventional
QN method is also suitable for tail fitting with hardware measurements of coarse time resolutions.
Thus, coefficients for the empirical error analysis with QN are derived equivalent to section 5.4.
The chapter concludes with a brief summary.
86
6.2. P ERFORMANCE O PTIMIZATION
for the linear case, so that the results can easily be obtained from summing terms.
We can describe a polynomial regression of order p as an approximation to a set of tail data
pairs (xi , qi ):
qi = a0 + a1 xi + . . . + ap xpi , i = {1, . . . , n} (6.1)
The number of n pairs can be arranged according to the Van-
dermonde [39] system of equations:
CDF(x)
1 x1 . . . xp1
a0 q1
1 x 2 . . . x p a 1 q2
2 q=Q(CDF(x))
.. .. = .. (6.2)
.. .. . .
. . . . . .
1 xn . . . xpn ap qn
Polynomial Regr.
σ̂err ,{a0 , . . . , ap }
When multiplied at the left by the transposed matrix we get:
P P p P store as σ̂err (n)
n x ... xi a0 qi
P P 2i P p+1 P
xi xi ... xi a1 x i qi n++
=
.. .. .. .. .. .. Regr. length
. . . . . P .p Q(n)≥0?
P p P p+1 P 2p no
xi xi ... xi ap x i qi
yes
(6.3)
This system of equations contains only summing terms which σ̂err,min =min{σ̂err (n)}
can easily be updated using recursions, as required by the
polynomial fitting stage. To obtain the regression coeffi- retrieve polynomial
coefficients
cients {a0 , . . . , ap } the matrix inverse must be calculated. The
symmetric arrangement of summing terms corresponds to a
calculate TJpp
Hankel matrix, which can be inverted very efficiently us-
ing the Levinson-Durbin algorithm. Here, an implementation
from [112] is used, which requires only 3p2 + 9p + 3 multiply F IGURE 6.2.: σ̂err based poly-
and divide operations, and is thus sufficiently fast for analyses nomial fitting.
up to the required polynomial of order p=4. The regression
error is calculated as standard deviation of a fitted polynomial with the tail data:
sP
n p 2
i=1 (qi − a0 − a1 xi − . . . − ap xi )
σ̂err (n) = (6.4)
n − (p + 1)
The Levinson-Durbin algorithm is included with the C/C++ simulation environment, where fast
statistical simulations are carried out. As with the sQN implementation, this allows for an in-depth
analysis of the algorithmic performance. Further, MATLAB is again used for post processing of
fitting results as well as for data representation.
In section 4.1.4 the computational effort of sQN and QN methods was already compared. With
the very efficient Levinson-Durbin recursion, the difference between QN and higher order poly-
nomials is marginal and thus, analyses of QP2, QP3 and QP4 methods are omitted here.
87
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
As already demonstrated with the simplified flow graph in figure 6.2, the minimum error can
be determined very quickly without amplitude scaling factor k. The minimum search is simply
carried out along the tail length n, where the error σ̂err is the goodness-of-fit measure.
When optimizing the performance, one likes to determine a best suited parameter configuration
for each of the polynomial fitting methods, equivalent to section 4.3.2. For the sQN method the
default configuration in equation (4.21) was derived as suitable interval for initial tail selection.
This configuration avoids outliers caused by statistical tail variations, and thus supports a robust
fitting behavior. For the polynomial methods the conservative parameters now have to be re-
specified.
Subsequent optimizations again use the median error Emed and interquartile range IQR of
TJpp estimates for performance analysis, since they are less influenced by outliers compared to
mean and standard deviation. In addition, the estimation loss EL was defined in equation (3.18)
to consider both error bias and spread. Further, higher order moments such as skewness ξ and
kurtosis κ are used for investigating the outlier behavior.
From the three conservative tail fitting parameters ∆Pt , ∆Tt and Pt,min , as defined and de-
scribed in section 4.3.2, ∆Tt exhibited poor performance, unless the Gaussian tail length was
known, which is usually not the case. Therefore, we only focus on the analysis of threshold Pt,min
and probability interval ∆Pt instead. Subsequently, performance optimizations are carried out
for each of the polynomial orders, starting with the first order or conventional Q-normalization
method.
88
6.2. P ERFORMANCE O PTIMIZATION
−1
0.1 10 0.2
0.05 0.1
|Emed|
−2
0.05
IQR
10
EL
0.02
0.01 0.02
−3
0.005 10 0.01
−3 −3 −3
10 10 10
1 1 1
−5 10 −5 10 −5 10
10 3 10 3 10 3
10 10 10
5 5 5
−7 10 −7 10 −7 10
Pt,min 10 ∆ Pt Pt,min 10 ∆P Pt,min 10 ∆ Pt
t
2
5 10
Skewness
Kurtosis
1
0 10
0
−5 10
−3 −3
10 10
1 1
−5 10 −5 10
10 3 10 3
10 10
5 5
−7 10 −7 10
Pt,min 10 ∆P Pt,min 10 ∆P
t t
(d) (e)
F IGURE 6.3.: First order polynomial regression (QN): median error Emed , interquartile range
IQR, estimation loss EL , skewness ξ and kurtosis κ surfaces over varying ∆Pt
and Pt,min . Test distribution: σRJ /ADJ =1/8, ADJ,uni =0.2 UI, σRJ =0.025 UI,
N =107 , K=250.
with a negative ξ, outliers are mostly located closer to the true timing budget, and are therefore
uncritical.
To summarize the properties of the QN method, best performance is achieved with the param-
eter product ∆Pt ·Pt,min chosen as large as possible. This avoids convergence failures as well
as negative skewed error distributions. In the subsequent performance comparison the same test
distributions will be used as for the analysis of the sQN method (see section 4.3.3). If the outlier
presence at small sample size N is neglected, the same ∆Pt values can thus be utilized.
Figure 6.4 shows the fitting performance of the second order polynomial regression (QP2), where
the worst case test distribution has again been identified at a jitter ratio σRJ /ADJ ≈1/8. The
performance metrics are the same as in figure 6.3.
Median error Emed , interquartile range IQR and estimation loss EL in the upper subfigures
behave quite similar to first order regression analysis. The magnitude of Emed in figure 6.4(a) is
significantly smaller than with first order polynomials, but this benefit is lost due to an increased
statistical spread. The overall estimation loss EL in figure 6.4(c) highlights an optimum parameter
region for the variable product ∆Pt ·Pt,min , forming a distinct valley. This is due to the IQR
course with a falling edge along the surface in figure 6.4(b), which already starts before the rising
edge of Emed in figure 6.4(a) influences the overall EL . Inside this region, best estimates are
obtained.
Skewness and kurtosis evidence a critical drawback for TJ estimates obtained from second or-
der polynomials. If only a small part of the distribution tail is selected for polynomial regression,
the resulting estimates will be highly scattered and contain a large amount of outliers. In fact,
89
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
−1
0.1 10 0.2
0.05 0.1
|Emed|
−2
0.05
IQR
10
EL
0.02
0.01 0.02
−3
0.005 10 0.01
−3 −3 −3
10 10 10
1 1 1
−5 10 −5 10 −5 10
10 3 10 3 10 3
10 10 10
5 5 5
−7 10 −7 10 −7 10
Pt,min 10 ∆ Pt Pt,min 10 ∆P Pt,min 10 ∆ Pt
t
2
5 10
Skewness
Kurtosis
1
0 10
0
−5 10
−3 −3
10 10
1 1
−5 10 −5 10
10 3 10 3
10 10
5 5
−7 10 −7 10
Pt,min 10 ∆P Pt,min 10 ∆P
t t
(d) (e)
F IGURE 6.4.: Second order polynomial regression (QP2): Emed , IQR, EL , ξ and κ surfaces
over varying ∆Pt and Pt,min . Test distribution: σRJ /ADJ =1/8, ADJ,uni =0.2 UI,
σRJ =0.025 UI, N =107 , K=250.
the positive skewness clearly indicates a heavy tail directing toward higher TJ estimates, and even
reaches an undesired peak for a worst case configuration. As will be shown later on with perfor-
mance evaluations, the median error Emed is negative for the selected test distribution. Thus, to
some degree one may tolerate positive skewed tails, but the magnitudes in figure 6.4(d) are by far
too large.
In order to ensure that second order polynomials achieve correct fitting results without outliers
and high skewness, a sufficiently large parameter product ∆Pt ·Pt,min must be selected. If the
threshold parameter Pt,min is not in use, the minimum probability corresponds to the granularity
of Pt,min =1/N =10−7 . The probability region for best tail fitting is then located at ∆Pt =5·105 ,
which is here used as default value for second order polynomials. It must still be adapted accord-
ingly for use with different sample sizes. Fortunately, the fitted tail part is not anymore restricted
to the linearized Q-tail, and may also include parts of the DJ component in order to achieve bet-
ter extrapolations. However, it becomes impossible to find an optimum parameter configuration
which is well suited for all test distribution shapes. Here we simply choose ∆Pt =5·10−3 ·N to
guarantee at least for a sufficiently large fitting region.
The same analysis from the previous paragraphs can also be carried out to investigate the estima-
tion performance of third and fourth order polynomials (QP3, QP4). Worst case test distributions
with uniform type DJ have again been identified at the jitter ratio σRJ /ADJ,uni ≈1/8.
With both polynomial orders, the regression error described by Emed , IQR and EL cannot
be influenced by the fitting parameter ∆Pt . Also the threshold parameter Pt,min provides best
performance only if reduced to its minimum Pt,min =1/N .
Unfortunately, the statistical random variation of tails has a highly misleading effect on higher
90
6.2. P ERFORMANCE O PTIMIZATION
−1
0.1 10 0.2
0.05 0.1
|Emed|
−2
0.05
IQR
10
EL
0.02
0.01 0.02
−3
0.005 10 0.01
−3 −3 −3
10 10 10
1 1 1
−5 10 −5 10 −5 10
10 3 10 3 10 3
10 10 10
5 5 5
−7 10 −7 10 −7 10
Pt,min 10 ∆ Pt Pt,min 10 ∆P Pt,min 10 ∆ Pt
t
2
5 10
Skewness
Kurtosis
1
0 10
0
−5 10
−3 −3
10 10
1 1
−5 10 −5 10
10 3 10 3
10 10
5 5
−7 10 −7 10
Pt,min 10 ∆P Pt,min 10 ∆P
t t
(d) (e)
F IGURE 6.5.: Third order polynomial regression (QP3): Emed , IQR, EL , ξ and κ plots for
∆Pt versus Pt,min . σRJ /ADJ =1/8, ADJ,uni =0.2 UI, σRJ =0.025 UI, N =107 ,
K=250.
−1
0.1 10 0.2
0.05 0.1
|Emed|
−2
0.05
IQR
10
EL
0.02
0.01 0.02
−3
0.005 10 0.01
−3 −3 −3
10 10 10
1 1 1
−5 10 −5 10 −5 10
10 3 10 3 10 3
10 10 10
5 5 5
−7 10 −7 10 −7 10
Pt,min 10 ∆ Pt Pt,min 10 ∆ Pt Pt,min 10 ∆P
t
2
5 10
Skewness
Kurtosis
1
0 10
0
−5 10
−3 −3
10 10
1 1
−5 10 −5 10
10 3 10 3
10 10
5 5
−7 10 −7 10
Pt,min 10 ∆ Pt Pt,min 10 ∆ Pt
(d) (e)
F IGURE 6.6.: Fourth order polynomial regression (QP4): Emed , IQR, EL , ξ and κ plots for
∆Pt versus Pt,min . σRJ /ADJ =1/8, ADJ,uni =0.2 UI, σRJ =0.025 UI, N =107 ,
K=250.
91
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
order polynomials, which generally impedes them to achieve accurate fitting results. In the case
of third order polynomials, skewness and kurtosis show heavy tails toward both sides (skewness is
close to zero while kurtosis is large). These heavy tails are caused by convergence failures of the
regression stage, where the fitted polynomial does not extrapolate the bathtub correctly down to
the target BER. This can for example be the result of a non-monotonic course of the fitted polyno-
mial, which contains a local minimum and/or maximum. Such a failure can be identified and dealt
by fitting a simple, strictly monotonic linear function into the distribution tail. However, this cor-
responds to a reduction of the regression order and thus, introduces a different error characteristic,
which finally leads to the heavy tails observed.
Fourth order polynomials highly suffer from random variations of measured bathtub tails. Al-
though skewness is almost zero and kurtosis reaches small values close to the normal distribution
case, the IQR plane shows an extremely large statistical spread. This highlights a general problem
of tail fitting algorithms with higher order polynomials. Instead of converging toward a simple
regression function, they tend to follow random data variations which mislead the extrapolation
result. In regression analysis this effect is commonly known as overfitting problem.
The ∆Pt parameter selection for third and fourth order polynomials is uncritical as it does in
fact not influence the estimation performance. We thus select a default value of ∆Pt =103 for both
polynomial orders.
The performance optimization of polynomial fitting methods concludes with a brief summary.
Each of the fitting methods has been investigated individually, in order to find an adapted, well
suited parameter configuration which drives the regression stage. These configurations are sum-
marized in table 6.1. Note that the parameters were selected using only uniform DJ type with
the worst case test distributions identified at σRJ /ADJ =1/8. They do not consider triangular or
quadratic curve shaped DJ. Thus, parameter configurations are primarily meant as compromise
solutions, where especially the higher order polynomials cannot be utilized over a broad variety of
test shapes. If for example, specific shapes are known, other configurations may be more suitable.
Polynomial ∆Pt
Order N = 104 N = 105 N = 106 N = 107 N = 108
1st 101 102 103 103 103
2nd 5 · 101 5 · 102 5 · 103 5 · 104 5 · 105
3rd 101 102 103 103 103
4th 101 102 103 103 103
TABLE 6.1.: Default parameter configuration of ∆Pt for polynomial tail fitting methods, derived
with worst case test distributions σRJ /ADJ =1/8 of uniform DJ type.
For N ≥106 the parameter ∆Pt is always constant besides second order polynomials, where
it must be chosen sufficiently large to avoid misleading outliers. By selecting ∆Pt =5·10−3 ·N ,
the optimum region from figure 6.4 is guaranteed, which has also been verified for different pa-
rameter surfaces over varying N . ∆Pt does not influence the performance of third and fourth
order polynomials, and was thus set to the default sQN values from equation (4.21). With first
order polynomials, parameter selection is more critical since the product ∆Pt ·Pmin must not ex-
ceed the minimum tail amplitude of test distributions. To guarantee this also for a small sample
size N ={104 , 105 }, the default sQN configuration is used as well. However, in this case the QN
method is also affected by outliers.
92
6.3. P ERFORMANCE A NALYSIS OF P OLYNOMIAL F ITTING M ETHODS
93
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
0.15
−1
0.1 10
med
EL
0.05
E
−2
10
−3
−0.05 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
F IGURE 6.7.: Sinusoidal type DJ: median error Emed (left) and estimation loss EL (right) for
polynomial tail fitting methods QN (a,b), QP2 (c,d), QP3 (e,f) and QP4 (g,h).
N ={104 , 105 , 106 , 107 , 108 }, K=250.
94
6.3. P ERFORMANCE A NALYSIS OF P OLYNOMIAL F ITTING M ETHODS
0.15
−1
0.1 10
med
EL
0.05
E
−2
10
−3
−0.05 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
F IGURE 6.8.: Uniform type DJ: median error Emed (left) and estimation loss EL (right) for
polynomial tail fitting methods QN (a,b), QP2 (c,d), QP3 (e,f) and QP4 (g,h).
N ={104 , 105 , 106 , 107 , 108 }, K=250.
95
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
0.15
−1
0.1 10
med
EL
0.05
E
−2
10
−3
−0.05 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
F IGURE 6.9.: Triangular type DJ: median error Emed (left) and estimation loss EL (right) for
polynomial tail fitting methods QN (a,b), QP2 (c,d), QP3 (e,f) and QP4 (g,h).
N ={104 , 105 , 106 , 107 , 108 }, K=250.
96
6.3. P ERFORMANCE A NALYSIS OF P OLYNOMIAL F ITTING M ETHODS
0.15
−1
0.1 10
med
EL
0.05
E
−2
10
−3
−0.05 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
0.15
−1
0.1 10
med
EL
0.05
E
−2
10
−3
−0.05 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.15
−1
0.1 10
med
EL
0.05
E
−2
10
−3
−0.05 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.15
−1
0.1 10
med
EL
0.05
E
−2
10
−3
−0.05 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
F IGURE 6.10.: Quadratic curve type DJ: median error Emed (left) and estimation loss EL (right)
for polynomial tail fitting methods QN (a,b), QP2 (c,d), QP3 (e,f) and QP4 (g,h).
N ={104 , 105 , 106 , 107 , 108 }, K=250.
97
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
98
6.4. C OMPARISON WITH S CALED Q- NORMALIZATION M ETHOD
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
99
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
100
6.4. C OMPARISON WITH S CALED Q- NORMALIZATION M ETHOD
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
101
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
102
6.4. C OMPARISON WITH S CALED Q- NORMALIZATION M ETHOD
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σ /A σRJ/ADJ
RJ DJ
0.1
−1
0.05 10
med
EL
0
E
−2
10
−0.05
−3
−0.1 10 −2 −1 0
−2 −1 0
10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ
103
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
0.12 0.12
QP2
QP2
0.09 0.09
QP4
QP4
EL
EL
0.06 0.06
QN QN
QP3
QP3
0.03 0.03
0.12 0.12
QP4
QP4
0.09 0.09
QP3
QN
EL
EL
0.06 0.06 QN
QP3
QP2
F IGURE 6.16.: Estimation loss EL of sQN, QN, QP2, QP3, QP4 fitting algorithms over varying
number of bins in the interval R=[32, 2048]. The figures evaluate the different
fitting algorithms with sinusoidal, uniform, triangular and quadratic curve DJ
types. N =108 , K=250, σRJ /ADJ =1/8, ADJ =0.2 UI.
indicates an over-fitting problem of higher order polynomials, caused by the statistical random
variation of tails.
104
6.5. E STIMATION E RROR A NALYSIS OF C ONVENTIONAL Q-N ORMALIZATION
TABLE 6.2.: Emed , IQR and EL coefficients for QN method, equation (6.6). ∆Pt /N =10−3
(large N ) and ∆Pt /N =10−2 (small N ), intervals specified above, K=250.
Therefore, only sQN and QN fitting methods are suitable for use with hardware jitter measure-
ments. In fact, higher order polynomials introduce too much error and can generally not guarantee
for pessimistic TJ estimation. With its fast algorithm, the QN method forms an alternative to the
computationally more expensive sQN method and thus, the error behavior is thoroughly analyzed
in the following section.
These coefficients are determined for each of the investigated test distributions and are subse-
quently listed in table 6.2. Estimation performance is investigated with respect to varying sam-
ple size N and bin density σRJ ·R. The intervals have been selected from original equidis-
tant grids of 75×51 nodes with σRJ ·R={0.8, . . . , 51.2} UI as well as N ={105 , . . . , 108 } and
N ={104 , . . . , 106 } respectively. Note, that the worst case shapes σRJ /ADJ for conventional QN
have been re-specified using figures 6.7-6.10, as they differ from the respective sQN values. A
constant ratio ∆Pt /N =10−3 for large N as well as ∆Pt /N =10−2 for small N again guarantee
for correct convergence of the algorithm. The r-squared statistic in the last column of the table
also highlights an excellent correlation.
The DNL model coefficients can be recalculated for the QN method as well, where a suitable
105
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
TABLE 6.3.: Emed , IQR and EL coefficients for QN method with included DNL error, equa-
tion (5.38). ∆Pt /N =10−3 (large N ) and ∆Pt /N =10−2 (small N ), K=200.
106
6.6. S UMMARY
The obtained empirical relations can now be used to design a jitter measurement system which
uses the conventional QN method for tail fitting, instead of the sQN method from chapter 5. There-
fore, the design equations can simply be reused, as they are also valid for QN. Equation (5.3) and
condition (5.10) guarantee for correct tail fitting if a minimum amplitude At,min is given. Fur-
ther, equation (5.20) and the selection chart in figure 5.9 specify a minimum RJ standard deviation
σt,min with known R, N and ∆Pt . Thus, QN and sQN errors may also be compared by simply
extending the design examples from section 5.5.
Example for Jitter Diagnostics The estimation performance resulting from R=128, N =107
and σRJ,min =7.57·10−3 , when inserted into equation (6.6) yields worst case errors (uniform DJ
type):
Emed = 2.0%, IQR = 2.3%, EL = 5.4% (sQN)
Emed = 2.9%, IQR = 2.4%, EL = 6.4% (QN)
For the given σRJ,min , the resulting coarse bin density is the major contributor to overall error and
thus, a rather small performance difference between QN and sQN is observed. This difference
becomes even smaller if additionally a DNL error of σDN L =0.05 UI is assumed and empirical
equation (6.7) is applied:
Emed = 2.3%, IQR = 3.4%, EL = 7.4% (sQN)
Emed = 3.2%, IQR = 3.0%, EL = 7.7% (QN)
which indicates that the statistical spread of the sQN algorithm is generally more affected by DNL
error than the QN method. However, this is also an effect caused by low bin densities, since DNL
has been modeled as a random step with time resolution 1/R (section 5.4.3) and can thus, also be
reduced with a higher time resolution.
Example for Production Tests The presented BIJM system can now also be used together
with the QN method, which only changes the empirical coefficients of equation (6.6) to the ones
specified in table 6.2. The results are depicted in figure 6.17. For pure Gaussian distributions the
error difference is marginal, because both QN and sQN methods behave similar. For sinusoidal
DJ, a performance difference is instead clearly visible.
0.5
0.4 QN
Sinusoidal DJ
0.3
Estimation Loss
0.2 QN
sQN
sQN
Only RJ
0.1
1 10 100
Number of Counters
F IGURE 6.17.: Estimation loss EL of sQN and QN methods over varying number of counters C,
with tt =20 ms, and worst case sinusoidal DJ (red) or pure RJ (black) case.
6.6. Summary
The performance of various recently proposed, polynomial tail fitting principles based on Gaussian
quantile normalization was compared. First a unifying optimization scheme (figure 6.1) equivalent
107
6. C OMPARISON OF G AUSSIAN TAIL F ITTING M ETHODS B ASED ON Q-N ORMALIZATION
to the one for the sQN method was developed, in order to realize all of the investigated polyno-
mial methods (QN, QP2, QP3, QP4). The required polynomial regression led to a Hankel matrix
notation which can be inverted very efficiently by using a Levinson-Durbin recursion from [112].
As a fundamental simplification compared to the sQN method, only the regression error from
equation (6.4) was shown to be applicable as fitness measure for tail fitting. Optimum parameter
regions for tail selection with ∆Pt were discussed for each of the polynomial orders. This led to
the selected parameter configurations in table 6.1.
Performance evaluations were first carried out for simulator environments (Rsim =3.3·105 ) with
respect to varying sample size N as well as DJ type (figures 6.7-6.10) for each of the polynomial
methods. The median error of QN with linear functions was shown to be strictly positive, which
is due to the asymptote in Q-domain, equivalent to the sQN method in section 4.2.1. Extrapolated
tails are thus always pessimistic. This beneficial property was generally not maintained with
higher order polynomials (QP2, QP3, QP4). Third and fourth order polynomials only produced
results with acceptable accuracy for a small range of jitter ratios. For RJ dominant shapes, they
were generally not able to correctly extrapolate tails, although forming simple linear functions in
Q-domain. This effect was shown to be a general over-fitting problem of higher order polynomials.
A comprehensive performance comparison with the sQN method based on ĉ1.2 with ∆Pt was
carried out in figures 6.11-6.15. The sQN method clearly achieved best performance for the cost
of a larger computational demand as already shown previously with figure 4.9. The QN method,
although less accurate, offers approximately 35 times faster fitting results. Further, it was shown
to possess the same positive error property as already described before and is thus, also well suited
for tail fitting.
An interesting effect was observed with second order polynomials for N ≥106 . The QP2 method
returned excellent results down to σRJ /ADJ ≈2·10−2 with triangular or quadratic curve shaped
DJ, where it even outperformed the sQN method. This is due to an optimum working region.
Nevertheless, the pessimistic extrapolation property of QN and sQN methods is not guaranteed
with the QP2 algorithm. Additionally, polynomials do not allow for recovering Gaussian model
parameters.
The influence of a limited time resolution with R bins on the extrapolation error was investigated
in figure 6.16. It again highlighted the effect of error oscillations at coarse resolutions, which was
also investigated in section 5.4.2. Results showed, that higher order polynomials are generally
more affected by such oscillations than sQN and QN methods. Further, the performance advantage
of the QP2 method with quadratic curve DJ was lost as soon as R<200.
Summarizing these results, only the sQN and QN fitting methods are well suited for use with
hardware jitter measurements. Higher order polynomials introduce too much error and can gen-
erally not guarantee for pessimistic tail extrapolations. Therefore, an error analysis for the QN
method was carried out in section 6.5, equivalent to the sQN error analysis in section 5.4. This
allowed to compare the extrapolation performance of both methods and thus, to choose the better
suited algorithm for a BIJM design. Results with the continued design examples from section 5.5
basically showed, that QN is generally less accurate than sQN, but is also less affected by differ-
ential non-linearity errors. Further, for pure Gaussian test distributions the performance difference
between QN and sQN methods becomes marginal.
Note that for a system design with the conventional QN method, all the important design equa-
tions from chapter 5 can be reused. This especially includes equation (5.3) and condition (5.10) to
guarantee for correct tail fitting results with respect to a minimum tail amplitude At,min , as well
as equation (5.20) and the selection chart in figure 5.9 to specify a minimum RJ standard deviation
σt,min . Parts of this chapter were also published in [C3,C7].
108
7. Jitter Analysis Method for Generalized
Gaussian Tail Extrapolation
In this chapter the Gaussian quantile normalization is brought into a generic analysis context.
Section 4.1.2 briefly mentioned the idea of generalizing the developed optimization scheme for
use with arbitrary tail shapes. As a possible application, especially the amplitude distributions of
high-speed signals may sometimes follow non-Gaussian shapes, including both heavy tailed distri-
butions as well as fast decaying tails. In such cases one likes to identify the unknown shape in order
to guarantee for correct signal recovery. As an extension to the pure Gaussian Q-normalization
thus, classes of tail distributions may be described with additional shape parameters. The gener-
alized principle then allows for correct tail extrapolation and jitter estimation of arbitrary tails as
long as they belong to the defined class.
Generalizing the scaled Q-normalization (sQN) method from chapter 4 with additional shape
parameters allows the linearizing principle to be extended to a complete class of tail distributions,
which includes the Gaussian function as a special case. Here, special attention is given to the
generalized Gaussian distribution (GGD) class of probability functions. It provides only one ad-
ditional shape parameter, and the goal is thus, to realize accurate tail extrapolations for GGDs.
The unknown shape parameter increases computational demand significantly, as it introduces an
additional degree of freedom to the optimization search space. The GGD class is also expected
to be less accurate than the sQN method for the special case of Gaussian tails. This is a natural
generalization effect when unknown tail characteristics and thus, less information is given to the
analysis system. On the other hand, the generalized method offers an outstanding opportunity,
it can identify underlying tail characteristics, and is thus able to tell whether a tail truly behaves
Gaussian-like or not.
In the following sections first the generalization principle is justified with a short literature
review, and criteria are described to ensure consistency with the existing RJ-DJ model as intro-
duced in section 2.3.1. Then various classes of generalized distributions and the corresponding
quantile normalization functions are presented. The generalized Gaussian distribution (GGD) is
described and its advantages compared to other function classes, before embedding it into the gen-
eralized optimization scheme for tail fitting. This scheme is implemented as efficient C++ routine.
Performance evaluations are carried out and compared against the sQN method and other fitting
principles. The chapter concludes with a brief summary.
109
7. J ITTER A NALYSIS M ETHOD FOR G ENERALIZED G AUSSIAN TAIL E XTRAPOLATION
0 UI 1 UI
"1" logic level
∆v PDFs
∆v
"0" logic level
∆t PDFs
∆t
F IGURE 7.1.: Eye diagram with timing jitter and amplitude noise PDFs [82].
Typical amplitude histograms in optical fiber channels follow a chi-squared distribution where the
critical tails decay very slowly and thus, cause bit errors at the receiver side. In order to min-
imize the BER, approaches based on maximum likelihood estimators and Viterbi-decoders are
utilized [2, 59, 126]. These approaches deal with non-linear noise properties of the transmission
medium, and require a careful analysis of observed amplitude distributions as well as the use of
generalized tail fitting methods.
Weinstein [135] was the first to propose a method for approximating the generalized Gaussian
distribution (GGD) class of functions. It extrapolates tails and estimates the BER without knowl-
edge of the underlying tail shape. Stojanovic [126] described an extrapolation method, which also
uses a subset of the GGD class. He also notes, that such tails especially appear in long haul optical
fibers and channels that suffer from severe signal distortions.
For short-range wireline communications, linear additive noise sources prevail. Thus, with
a large number of random processes involved, amplitude noise and timing jitter PDFs mostly
follow a Gaussian tail. Nevertheless, in [75, 104] different jitter types are classified and the RJ
section also defines an arbitrary non-Gaussian case which cannot be decomposed by conventional
fitting methods. However, non-Gaussian timing jitter has so far only been described for soliton
transmission, as a result of the Gordon-Haus effect [37, 44].
When focusing on the analysis of non-Gaussian tails thus, amplitude noise clearly dominates
the practical use case. Fortunately, with the definition of the unit interval (UI), fitting methods
are not restricted to the analysis of timing jitter and can equivalently also be applied to amplitude
histograms. In this context, an accurate jitter decomposition method for arbitrary RJ shapes is
subsequently developed. It is consistent with the existing RJ-DJ model [104] and hence, forms a
logical extension to the commonly accepted modeling approach.
The method can easily be derived from the generalized sQN scheme in figure 4.6. Therefore,
only a suitable quantile function for the generic normalization stage must be determined, as de-
picted in figure 7.2. Q(p) forms the heart piece for tail linearization. It may normalize a specific
tail shape, or a whole class of distributions when described by one or more shape parameters (α,
β, . . .). Note that every additional parameter introduces another degree of freedom for the opti-
110
7.1. I NTRODUCTION TO G ENERALIZED TAIL F ITTING
Generalized q
Γ(1/α) 1
R x u−1 −t
β=σ γ(u, x) = Γ(u) 0 t e dt
Gaussian Γ(3/α)
R∞
Γ(u) = 0 e−t tu−1 dt sgn(. . .) ... sign function
x ∈ [0, ∞) p ∈ [0, 1]
x
Exponential 1 −σ − ln(1 − p)
σe
(
Generalized −1− 1 − ln(1h− p) if α = 0
1 x−µ α
1+α·
i
Pareto σ σ −1/α 1 − (1 − p)−α if α 6= 0
Generalized 1 1 1
−α
(
− ln − ln(p)
if α = 0
Extreme σ (1 + αz)−1− α · e−(1+αz) h −α i
Value z = (x − µ)/σ −1/α 1 − − ln(p) if α 6= 0
TABLE 7.1.: Quantile normalization functions for different tail distributions [106].
mization scheme and thus, significantly increases computational demand. Once the tail parameters
have been identified, the timing budget TJpp can easily be determined at the target BER=10−12
with:
TJpp = tL + tR (7.1a)
−12
tL = µL + σL · Q(10 /AL , α, β, . . .) (7.1b)
tR = µR + σR · Q(10−12 /AR , α, β, . . .) (7.1c)
which corresponds to the inverse probability function for an expected tail shape of unit amplitude,
unit standard deviation and zero mean.
The quantile normalization Q(p) may utilize additional shape variables to include the tail char-
acteristics of entire function classes. In table 7.1 Q(p) is listed for various candidates. The Gaus-
sian and generalized Gaussian functions are defined over the complete real axis. That is, they
are symmetric for positive and negative values of x. Other distributions such as the exponential,
generalized Pareto or generalized extreme value PDFs are bounded toward negative values of x.
The PDF definitions use a positive range for x to denote the right sided normalization function,
which is applied to positive bathtub tails with p=F (x). If measured tails belong to the domain
of attraction of the normalizing function, the optimization scheme from figure 7.2 is thus able to
identify a best suited set of tail parameters.
The presented distribution classes give an outline to possible tail characteristics, which may
exhibit power-law, exponential or Gaussian-like behavior instead of the pure Gaussian case. The
111
7. J ITTER A NALYSIS M ETHOD FOR G ENERALIZED G AUSSIAN TAIL E XTRAPOLATION
0.8
α=1.0
Shape α Distribution
0.6
→0 Dirac impulse
0.5 Gamma
PDF
α=2.0
0.4
α=100 1.0 Laplace
0.2 2.0 Gaussian
→∞ Uniform
0
−5 0 5
σ
correct choice for an expected tail shape is crucial as it highly influences extrapolation results.
Generalized extreme value (GEV) distributions and generalized Pareto (GP) distributions are com-
monly used in extreme value theory, and their use for nonparametric tail extrapolation and thresh-
old models in general has been discussed extensively in literature [19, 115, 117]. Nevertheless,
here we focus on the generalized Gaussian distribution (GGD), which offers two major advan-
tages compared to other generalizations. First, it represents a simple and direct generalization
of the pure Gaussian function. With only one additional shape parameter α, it includes impor-
tant special cases such as the Gaussian (α=2) or the exponential (α=1) tails. Second, it is fully
consistent with the commonly accepted RJ-DJ model from [104], and thus allows for decompos-
ing a total distribution into its non-Gaussian random and bounded deterministic components. In
fact, the additional shape parameter α extends the existing model with the ability of tail shape
characterization.
α −| x−µ |α
PDF(x) = f (x, α, β, µ) = e β (7.3a)
2 · β · Γ(1/α)
s Z ∞
Γ(1/α)
β=σ , Γ(u) = e−t tu−1 dt (7.3b)
Γ(3/α) 0
where [−∞ < x < ∞] and α > 0. Γ(u) is the gamma function, which is required for amplitude
normalization so that α can be varied independently. Note that β is a dependent scale parameter,
directly related with σ and α, and hence, it does not influence the tail shape.
The advantages of this representation form are, that only the α parameter defines the tail shape
and that the GGD class covers several special function types. With α=2 the normal distribution is
obtained, while for α=1 the Laplace distribution with exponential tails, and for α=0.5 the heavy
tailed Gamma distribution. A small shape value yields an impulsive function with slowly decaying
tails, while a large value leads toward the uniform distribution. This behavior is also depicted in
figure 7.3. The α parameter thus offers a flexible way for representing a generic distribution class
with exponential tail behavior.
The quantile normalization for GGD functions is derived as inverse of the CDF, which is ob-
112
7.2. I MPLEMENTATION OF A LGORITHM
where γ(u, x) is the incomplete gamma function and sgn(. . .) the sign function. The inverse CDF
can be written as:
1/α
F −1 (p, α, σ, µ) = γ −1 1/α, |2p − 1| · β · sgn(2p − 1) + µ (7.5)
For the quantile normalization Q(p, α)=F −1 (p, µ=0, σ=1) the case with unit standard deviation
and zero mean must be considered, and thus:
s
1/α Γ(1/α)
Q(p, α) = γ −1 1/α, |2p − 1| · · sgn(2p − 1) (7.6)
Γ(3/α)
The resulting equation describes a transform which allows for the linearization of arbitrary GGD
tails. When inserted into the optimization scheme from figure 7.2 we have four unknown variables:
α, A, σ and µ. This four-dimensional search space has to be dealt by an efficient search algorithm,
as described in the subsequent section.
Equivalent to the analysis of test distributions with Gaussian tails in section 3.2.2 GGD, test
distributions must also be created in order to analyze and compare the estimation performance
of tail fitting algorithms. Therefore, the composition principle with RJ and DJ can be reused by
replacing the Gaussian RJ generator with a GGD jitter source. The additional shape parameter is
denoted as αRJ , and all the prior performance metrics such as estimation loss EL or median error
Emed can be reused.
An important issue relates to correct GGD jitter generation. Since the quantile normalization
function describes a probability p∈[0, 1] in terms of the amplitude of a normalized random vari-
able, one can use equation (7.5) for random sample generation. Figure 7.4 demonstrates this
principle, where a uniform random process Juni generates jitter samples which are used as input
to the GGD normalization function. The required standard deviation σ is obtained by scaling,
while a non zero mean µ yields additional data offset.
Juni Jggd
Q(t, α) · σ + µ
0 1
113
7. J ITTER A NALYSIS M ETHOD FOR G ENERALIZED G AUSSIAN TAIL E XTRAPOLATION
million random samples, they must be generated very quickly. Therefore, this section especially
focuses on speed optimization for the GGD jitter source and the optimization scheme in figure 7.2,
which faces the four-dimensional search space.
In [26] the γ −1 function has been realized very efficiently with a third-order Schröder iteration,
supported by the Newton-Raphson method. A target accuracy of rel =10−6 is typically reached
after the second or third iteration step. Also the included complete gamma (7.3b) and incom-
plete gamma (7.4b) functions are realized with minimax rational approximations and a uniform
asymptotic expansion respectively. This yields an excellent computational efficiency.
The described functions have been realized as C++ routines, where one million γ −1 function
calls require approximately three seconds of simulation time on an Intel Core Duo 2.2GHz laptop.
This is still not fast enough, neither for tail parameter search with the scheme in figure 7.2, nor for
the GGD random generator in figure 7.4. Especially when gathering millions of random samples,
test distributions have to be generated significantly faster. Therefore, additional minimax approx-
imations of Q(p, α) have been realized, which support certain discrete α values and achieve an
additional speed up of more than one order of magnitude. As selectable grid values, they cover
shapes in a range from α=[1, 10] with logarithmically scaled distance and a maximum relative er-
ror of rel =10−6 . These minimax approximations allow for speeding up the optimization process
with a fast initial search grid and thus, also achieve a quick generation of random values.
An efficient realization of the optimization scheme in figure 7.2 requires a fast identification
of the global error minimum. The regression stage fits a simple linear function to the tail part of
n outermost tail samples by reusing the recursions (4.5) of the sQN method from section 4.1.1.
The regression is thus already optimized with respect to fast line slope and offset recovery. The
regression error σ̂err is the mean square error of fitted data pairs (qi , xi ):
s
Pn 2
i=1 qi − o − s · xi
σ̂err = (7.7)
n−2
The two remaining parameters scaling factor k and shape α obviously define the quantiles, here
given as qi values. Both parameters have to be identified in the context of an additional outer
optimization, characterized by another two dimensional minimum search.
In section 4.3 the ratio T̂ =σ̂err /s was also introduced as a fitness measure, suggested by
Scholz [117] for judging the appropriateness of a fitted line. For GGD tail extrapolation, this
criterion now drives the outer optimization. The optimization process is composed of two steps.
An initial search grid identifies a coarse error minimum, which is then refined by a search routine.
The initial search grid locates the global minimum, where the tail length n is pushed in as far as
possible. This corresponds to the same algorithmic principle as already described in section 4.3.
For each of the (k, α) grid values the T̂ (n) minimum is determined as a function of tail length
n, and the grid value with maximum tail length is selected. The search grid is logarithmically
scaled with a distance factor of ∆k=1.2 for both k and α variables. The investigated intervals
are k=[101 , 103 ] and α=[1, 101 ], while ∆Pt ≥102 is used as minimum tail interval for outlier sup-
pression. The resulting two-dimensional surface is thus a function of k and α, and typically forms
a narrow valley where the global minimum is hardly distinguishable along the bottom course. A
typical example is given in figure 7.5.
The second optimization step is a local refinement of the resulting plane, carried out with a
bounded search algorithm. Starting with the initial grid pair (kin , αin ), the bounds are:
αlo ≤ αin ≤ αup , αlo = max(1, αin /∆k), αup = min(10, αin · ∆k) (7.8)
3
klo ≤ kin ≤ kup , klo = max(1, kαlo ), kup = min(10 , kαup )
where kαlo and kαup are determined from the search grid as k-minimums which appear at the
αlo and αup values respectively. The refinement stage finally determines the minimum values
114
7.3. P ERFORMANCE A NALYSIS
−2
10
−3
10
T
−4
10
2
10
2
10
1
10 1
10
0 0
k 10 10 α
F IGURE 7.5.: T̂ surface for two-dimensional minimum search with initial search grid.
(kopt , αopt ) on the parameter surface. It uses the BOBYQA algorithm [113], which showed good
performance and a fast convergence compared to other search algorithms from the NLopt [67]
nonlinear optimization library. As convergence criteria a minimum relative parameter variation of
xtol ≤10−5 or a maximum function count of three hundred iterations are used.
The overall optimization converges sufficiently fast, especially due to the minimax function
approximations of the initial search grid. The refinement step typically requires 30-40 function
calls. As with the sQN method in section 4.1.4, computational demand again depends on the time
resolution R, since the algorithm has to process all distribution bins. For Rsim =3.3·105 and the
same test distribution as for the performance analysis in figure 4.9 (ADJ =0.2 UI, σRJ =0.05 UI,
αRJ =2), the 2.2GHz laptop typically requires approximately one minute, which is acceptable for
model simulations. In order to speed up the optimization in subsequent analyses, the simulator
time resolution is reduced to some degree (Rsim =105 ) without significant influence on the error.
115
7. J ITTER A NALYSIS M ETHOD FOR G ENERALIZED G AUSSIAN TAIL E XTRAPOLATION
−1 0
0.015 10 10
0.01
−1
0.005 10
med,α
Emed
−2
L
0 10
E
−2
−0.005 10
−0.01
−3 −3
−0.015 10 −2 −1 0
10 −2 −1 0
−2 −1 0
10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
(a) Emed , αRJ ={1.0, 1.44, 2.0} (b) EL , αRJ ={1.0, 1.44, 2.0} (c) Emed,α , αRJ ={1.0, 1.44, 2.0}
−1 0
0.015 10 10
0.01
−1
0.005 10
med,α
Emed
−2
L
0 10
E
E
−2
−0.005 10
−0.01
−3 −3
−0.015 10 −2 −1 0
10 −2 −1 0
−2 −1 0
10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
(d) Emed , αRJ ={3.0, 6.2, 8.9} (e) EL , αRJ ={3.0, 6.2, 8.9} (f) Emed,α , αRJ ={3.0, 6.2, 8.9}
αRJ =1.0 αRJ =1.44 αRJ =2.0 αRJ =3.0 αRJ =6.2 αRJ =8.9
α0 =1.0 α0 =1.44 α0 =2.0 α0 =3.0 α0 =6.2 α0 =8.9
F IGURE 7.6.: Median error Emed , estimation loss EL , and shape error Emed,α over varying test
distribution shape: N =107 , ∆Pt =102 , uniform DJ, Rsim =105 , K=250.
With Rsim =105 and a distribution sample size of N =107 , tails can be fitted sufficiently accurate
and correspond to an extrapolation which ranges over five orders of magnitude.
In figure 7.6 the estimation performance of the GGD tail fitting method is investigated with
respect to varying jitter ratio σRJ /ADJ and αRJ . The solid lines with filled markers denote the
proposed GGD method, while dashed lines show the results obtained when αRJ is already known
to the search algorithm. In this case the constant α0 =αRJ replaces the shape variable and yields a
reduced search space for the fitting algorithm. The special case of α0 =αRJ =2 for example, gives
the simplified form of Gaussian tail fitting with scaled Q-normalization from chapter 4.
At αRJ =1, the error bias Emed tends to be negative and thus, to slightly underestimate the TJ
values. This is not the case for known α0 =αRJ , which is again due to the asymptotic tail behavior
in quantile domain. For larger αRJ values, the median error of the GGD method also becomes
positive, which is an effect caused by very steep tails.
Over all the investigated distribution shapes, EL yields a worst case error which is less than
2.5%. If αRJ >3, the estimation loss also shows that the generalized method can directly compete
with the known α0 scenario. Unfortunately, this is due to highly overestimated shape values, as can
be seen in figures 7.6(c) and 7.6(f). Here, the median shape error Emed,α is calculated equivalent
to Emed , using estimated shape values:
Especially with steep tails, either realized by small jitter ratios σRJ /ADJ or a large αRJ , the
algorithm rather overestimates the true shape instead of correctly fitting a small tail amplitude.
Therefore, an exact tail shape identification of distributions can only be carried out in a very
limited sense.
116
7.3. P ERFORMANCE A NALYSIS
−1 0
0.04 10 10
0.02
−1
10
med,α
Emed
−2
L
0 10
E
−2
10
−0.02
−3 −3
−0.04 10 −2 −1 0
10 −2 −1 0
−2 −1 0
10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
F IGURE 7.7.: Emed , EL , Emed,α for various distribution shapes equivalent to figure 7.6, with
constant shape parameter αRJ =2.0 (Gaussian case) and varying sample size
N ={107 , 108 , 109 }, K=250.
−1 0
0.04 10 10
0.02
−1
10
med,α
Emed
−2
L
0 10
E
E
−2
10
−0.02
−3 −3
−0.04 10 −2 −1 0
10 −2 −1 0
−2 −1 0
10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
F IGURE 7.8.: Emed , EL , Emed,α for various distribution shapes equivalent to figure 7.6, with
constant shape parameter αRJ =2.0 (Gaussian tails) and varying time resolution
R={105 , 1024, 128}, N =107 , K=250.
Acceptable results are obtained with αRJ ≤2 and large jitter ratios σRJ /ADJ ≥0.5 (RJ dominant
case). A possible way to extend this application range is to increase the sample size N , which has
been kept comparable small so far. Therefore, in figure 7.7 the αRJ =2.0 case is again investigated
with respect to varying N . Although a significant improvement can be achieved for both Emed
and EL , αRJ still remains highly overestimated. Also note, that an exponential increase of the
sample size can soon lead to unacceptable simulation or measurement times.
117
7. J ITTER A NALYSIS M ETHOD FOR G ENERALIZED G AUSSIAN TAIL E XTRAPOLATION
−1 −1 −1
10 10 10
EL
EL
E
−2 −2 −2
10 10 10
−3 −3 −3
10 −2 −1 0
10 −2 −1 0
10 −2 −1 0
10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
−1 −1 −1
10 10 10
EL
EL
E
−2 −2 −2
10 10 10
−3 −3 −3
10 −2 −1 0
10 −2 −1 0
10 −2 −1 0
10 10 10 10 10 10 10 10 10
σRJ/ADJ σRJ/ADJ σRJ/ADJ
F IGURE 7.9.: Estimation loss EL over varying jitter ratio σRJ /ADJ . The test distributions
are the same as in figure 7.6, with αRJ ={1.0, 1.2, 1.44, 1.73.2.0, 3.0}, N =107 ,
Rsim =105 and K=250.
the scenario with constant and known α0 (dashed lines), the generalized fitting method performs
significantly worse. Especially with R=128 a large peak toward negative Emed is noticed. This
behavior is highly undesired as it leads to optimistic TJ estimates. The error peak ranges from a
large jump of shape estimates in figure 7.8(c), which also evidences the critical jitter ratio limit
where the GGD method starts to fail.
118
7.4. S UMMARY
and underestimates it at small ones. This makes the method only applicable down to a certain
limit. The error EL is quite large due to the approximation of GGDs in double-log domain, but
the method is especially simple and thus, very fast.
The performance comparison with the sQN method highlights a possible application of the
generalized method, which is for tail shapes in the interval range αRJ =[1, 2]. Here, the sQN
method becomes highly optimistic with negative error bias, because it always assumes Gaussian
tail behavior. The test distributions instead are heavy tailed and thus, follow a flat course which
cannot be tracked correctly by sQN. If σRJ /ADJ is sufficiently large and the TJ error must be
strictly positive, Weinstein’s method is better suited, because it always guarantees for pessimistic
tail extrapolations.
7.4. Summary
A generic optimization scheme for non-Gaussian tail fitting was presented, which led to a jitter
analysis method for generalized Gaussian distributions (GGDs). Therefore, the Gaussian quan-
tile normalization principle from chapter 4 was reused. The method is fully consistent with the
existing RJ-DJ model and utilizes only one additional shape parameter to describe different tail
characteristics.
In simulators the proposed method correctly fits and extrapolates tails that belong to the GGD
class of functions, although clearly outperformed by the case of known shape parameter α. The
major drawback of the generalized method results from the optimistic extrapolation property. This
is due to a general overestimation of shape values, especially at α=2 and α=3, as shown in fig-
ure 7.6. This problem can only be dealt by a significant increase of the sample size N (figure 7.7).
In hardware scenarios the limited time resolution R introduces additional error as well as a lower
bound for analyzable jitter ratios σRJ /ADJ (figure 7.8). Both effects further augment the opti-
mistic error nature of the GGD method.
A comparison of different methods was also carried out in figure 7.9. It clearly shows the
performance advantage of the generalized method with heavy tailed distributions (αRJ <2) in
comparison with the sQN method from chapter 4. If a pessimistic tail extrapolation with αRJ ≥1
is the crucial design criterion, Weinstein’s method from [135] is the best suited choice.
The proposed method has been implemented using a minimum of constraints in order to facili-
tate a broad application field. Improvements can possibly be realized by considering only specific
shapes or narrow parameter ranges, and by combining them with more suitable optimization cri-
teria and fitness measures then the ones utilized here.
Summarizing these results, the GGD method must be utilized very carefully when characteriz-
ing unknown tail shapes. In simulation scenarios it provides accurate extrapolations, although with
an undesired optimistic error bias. A reduction of the number of bins R as required for hardware
applications is not recommended, since obtained results become highly unreliable. Summarized
parts of this chapter have also been published in [C5].
119
8. An Accurate Behavioral Model for
High-Speed PLLs
To highlight the practical aspect of jitter analysis methods, a typical application with system mod-
eling and simulation is presented. Therefore, an accurate behavioral model of a high-speed trans-
ceiver is implemented. It has also been realized as test structure [62], and the goal is to analyze
its jitter behavior as typically required for system development and verification. The transceiver is
modeled as system level model according to the top-down methodology [63], using SystemC as a
C++ based library [38, 41, 61]. Therefore, system blocks are first brought into an abstract design,
and then successively refined down to the desired accuracy.
The model realizes a charge-pump PLL (CPLL) for high-speed clock and data recovery at
3Gb/s, and is intended for the S-ATA communication standard [118]. It uses accurate transient
simulations to analyze and predict the jitter behavior in terms of phase noise PSDs and jitter trans-
fer functions. As an enhanced version of a prior event driven approach [50], it affords accurate
analysis of jitter generation and propagation effects inside a CPLL.
In the following, the basic CPLL model is described together with prior modeling approaches
before proceeding to an enhanced event-driven model for accurate behavioral simulations. In the
analysis section, phase noise spectra, time domain parameters and jitter transfer characteristics are
derived and compared with measurements from the test structure, to demonstrate applicability of
the model, as well as the tail fitting method from chapter 4.
Icp
data in
up
LF gr VCO
ip VO output
BB−PD fc
R0 C1
down
VC1
Icp C0
VC0
121
8. A N ACCURATE B EHAVIORAL M ODEL FOR H IGH -S PEED PLL S
Only transient simulations are able to accurately reflect the true time domain behavior and cope
with the non-linear loop dynamics. For such simulations, two different modeling principles have
been developed in recent years. The first one is based on event-driven concepts [1, 21, 46, 50, 77],
where the analog part of the CPLL (i.e. the loop filter) is replaced by a set of non-linear recursive
equations. These are exact solutions to the loop filter difference equations. Hence, they are able
to determine the exact time instant of the subsequent VCO clock, which transforms the analog
circuit part into an event-driven block. Since all other circuit components are digital blocks, the
complete CPLL can thus be simulated as an event-driven system, which is typically characterized
by its non-uniform simulation time steps.
The second principle [107,108] instead, uses uniform time steps for simulation. The analog loop
filter is initially converted into discrete-time via impulse-invariance or bilinear transform. The
quantization noise of asynchronous events is subsequently considered by varying the amplitude
of digital signals according to the location of the transition edge between sampling instances.
This yields a highly accurate signal representation in discrete-time, combined with a fast model
implementation which has also found use in modern PLL simulators [109].
In this section an enhanced event-driven model is implemented according to the first model-
ing technique as explained in [46, 50]. Although the principle from [107] might afford very fast
simulations and thus be an intuitive choice, it does not allow for a dynamic variation of model
components. This is an important point, since the present CPLL model is only a part of an overall
transceiver structure which requires a careful stability analysis. Especially when changing be-
tween different operating modes, such as startup phase and normal lock-to-data operation, a stable
PLL behavior must always be guaranteed. The event-driven model uses the actual physical state
of analog filter components to recursively determine successive states. Thus, it directly reflects
the physical behavior of the PLL loop filter at any calculated time instant, which also allows for
dynamic filter variations. This way, one can easily switch between different operating modes. A
discrete-time filter instead cannot include dynamic variations, because its state variables do not
reflect the actual physical state.
As modeling environment, the SystemC [41, 61] programming language affords simple system
level simulations, and can directly embed the developed jitter analysis methods. In the following
subsections a closer look to the event-driven model is given. The basic model is based on an exact
solution for the 3rd order CPLL [50], and is here enhanced with a noise model for the VCO [122],
an additional parasitic pole for the gain regulator, and some BB-PD caused non-ideal effects.
122
8.1. M ODELING P RINCIPLE
where Kv is the linear frequency slope and f0 the zero voltage frequency. In a practical VCO, this
simple relation is hardly valid for the complete tuning curve of the oscillator. However, in lock-to-
data mode it is constantly driven at the transmission rate and thus, must only be valid for a local
operating point. The oscillator output phase can be expressed as integral of the VCO frequency,
and hence: Z t0 +t
Θvco (t) = Θvco (t0 ) + 2π [Kv VO (τ ) + f0 ] dτ (8.2)
t0
The charge-pump generates discrete current pulses for a second order loop filter. These current
pulses are assumed as constant ip ∈ {+Icp , 0, −Icp }, without depending on the loop filter voltage
or other non-linear effects. The behavior can thus be described in time domain with two differential
equations, obtained by Kirchhoff’s laws:
di1 (t) C0 + C1 ip
+ i1 (t) · = (8.3a)
dt C0 R0 C1 C0 R0
di0 (t) C0 + C1 ip
+ i0 (t) · = (8.3b)
dt C0 R0 C1 C1 R0
Solving these differential equations and representing the variables in voltage domain, leads to the
recursive equations:
1 1 i0 (t0 ) − β2
Z
−β1 t
VC0 (t) = · i0 (τ )dτ = VC0 (t0 ) + · 1−e + β2 · t (8.4a)
C0 C0 β1
1 1 β2 − i0 (t0 )
Z
VC1 (t) = · i1 (τ )dτ = VC1 (t0 ) + · 1 − e−β1 t + β3 · t (8.4b)
C1 C1 β1
with
VC1 (t0 ) − VC0 (t0 )
VO (t) = VC1 (t), i0 (t0 ) =
R0
C0 + C1 C0 · ip C1 · ip
β1 = , β2 = , β3 =
C0 R0 C1 C0 + C1 C0 + C1
where the gain regulator with parasitic pole fc and gain gr in figure 8.1 is not considered for
the moment. We can now insert equation (8.4b) into equation (8.2), and yield the final recursive
equation for the VCO phase:
( " #)
K i (t ) − β t 2
v 0 0 2
Θvco (t) = Θvco (t0 ) + 2π t f0 + Kv VO (t0 ) + 1 − e−β1 t − β1 t + β3
C1 β12 2
(8.5)
The goal is to determine the exact time instant t where the VCO phase reaches Θvco (t)=2π and
thus, produces the next clock edge for the BB-PD, which finally closes the feedback loop. Since
equation (8.5) cannot be inverted, a Newton iteration is utilized to recursively identify t. In our
case the function of interest is Θvco (t) which converges toward 2π and thus, we have the iteration:
The Newton method typically converges after the second or third iteration step by reaching the
maximum simulator time resolution of 1fs. With this basic model from [50], additional non-ideal
and non-linear effects can now be included, in order to significantly improve accuracy of the given
CPLL model.
123
8. A N ACCURATE B EHAVIORAL M ODEL FOR H IGH -S PEED PLL S
g0
√
N (0, f s)
fjit
g1 1 − z −1
fs
flicker
g2 filters
depicted in figure 8.2, where the gain factors are calculated as follows:
p
g0 = f1 10(A1 /10) , ∼ 1/f 2 noise (8.8a)
(∆/20) 2
g1 = 10 ∆ = AP hN − 10 log10 (4π
, /fs2 ) ,
∼ 1/f noise 0
(8.8b)
g2 = f1 ff l · 10(A1 /20−1.2704) , ∼ 1/f noise 3
p
(8.8c)
The flicker noise filter bank according to [122] is realized with eight filters of increasing cut-off
frequency, where the first one is located at 100 Hz.
yk [n] = ak yk [n − 1] + bk g2 x[n] , k = [0, . . . , 7] (8.9a)
10(k+2) 10(k/2+2)
ak = 1 − 2π , bk = 2π (8.9b)
fs fs
X
ff l [n] = yk [n] (8.9c)
where x[n] are the random samples from the noise generator at the rate fs . The overall jitter
frequency fjit is the sum of the three noise components:
fphn [n] = g1 (x[n] − x[n − 1]) , ff 1 [n] = g0 x[n] (8.10a)
fjit = ff l + fphn + ff 1 (8.10b)
The sampling rate fs must be sufficiently large to guarantee for correct noise generation and thus,
simply the maximum VCO frequency is used.
124
8.1. M ODELING P RINCIPLE
−3
x 10
2
1 VC0
Voltage [V]
−1 VO
−2 VC1
−3
−4 −6
x 10
5
CP
0
I
−5
0 0.2 0.4 0.6 0.8 1 1.2 1.4
time [s] −8
x 10
F IGURE 8.3.: Loop filter voltage behavior depending on input current Icp , the voltages can be
found in the block scheme, figure 8.1.
"(
gr Kv β2 − i0 (t0 )
Θvco (t) = Θvco (t0 ) + 2π f0 t + · 1 − e−ωc t − ωc t − · · ·
C1 ωc ωc − β1
! !
ωc2 β3
− 2 1 − e−β1 t − β1 t + − VC1 (t0 )C1 1 − e−ωc t − ωc t + · · ·
β1 ωc
# )
β3 ωc 2 Kv
+ t + VO (t0 ) 1 − e−ωc t (8.12)
2 ωc
The VCO control voltage VO is now given by the gain regulator output which introduces an addi-
tional, third state variable. The subsequent VCO clock period is calculated using equation (8.12)
and replaces equation (8.5) of the prior basic model.
Figure 8.3 shows the difference of the voltage behavior between basic and enhanced model,
characterized by VC1 and VO respectively. The additional parasitic pole highlights a significant
influence on the voltage course, and is thus essential for reflecting the real circuit behavior.
The digital BB-PD block is modeled with a propagation delay tdel and the non-ideal behavior
of data latches. When the VCO clock triggers close to an input data transition, digital output data
125
8. A N ACCURATE B EHAVIORAL M ODEL FOR H IGH -S PEED PLL S
suffers from meta-stability, offset voltages, or setup and hold time violations. These effects are
considered by introducing a random output value at the logical up or down signals, if one of the
criteria is violated.
126
8.2. J ITTER AND P HASE N OISE A NALYSIS
−70 −70
−80 −80
−100 −100
−110 −110
−120 −120
−130 −130
−140 −140
2 4 6 8 10 12 14 2 4 6 8 10 12 14
Frequency [Hz] 7 Frequency [Hz] 7
x 10 x 10
−70 −70
−80 −80
Phase Noise PSD [dBC/Hz]
−100 −100
−110 −110
−120 −120
−130 −130
−140 −140
2 4 6 8 10 12 14 2 4 6 8 10 12 14
Frequency [Hz] 7 Frequency [Hz] 7
x 10 x 10
F IGURE 8.4.: Measured (dashed red) and simulated (solid blue) phase noise PSD over different
parameter settings.
signals, additional spectral spurs are observed. They originate from a 50MHz clock which is used
as on-chip reference frequency. Besides these spurs, the model is able to accurately reflect the true
phase noise behavior. In combination with fast simulations, this allows for a deep and thorough
system exploration.
Several common time-domain parameters for jitter characterization can also be determined from
the statistics of absolute jitter. They form simple alternatives to the spectral description of phase
noise. Typically, one specifies the standard deviation or RMS value of accumulated jitter:
(m)
σacc (m) = RMS jacc = RMS jabs,k − jabs,k+m (8.13)
where jabs,k (section 2.1.2) is the absolute jitter sequence of the VCO clock. It can be calculated
from the autocorrelation function rjabs of absolute jitter values [42, 99]:
2
σacc (m) = 2 · (rjabs (0) − rjabs (m)) (8.14)
Commonly used RMS values include absolute jitter σabs , long term jitter σlt , period jitter σper or
maximum jitter σmax [22, 42], and are defined as:
q
σabs = RMS jabs,k = rjabs (0) (8.15a)
√
σlt = σacc (m) = 2 · σabs (8.15b)
m→∞
σper = σacc (1) (8.15c)
σmax = max σacc (m) (8.15d)
where especially σlt is often used to specify the overall performance of a PLL. In figure 8.5 the
RMS value of accumulated jitter σacc (m) is depicted for the same parameter configurations as in
127
8. A N ACCURATE B EHAVIORAL M ODEL FOR H IGH -S PEED PLL S
0.06 0.12
0.05 0.1
0.04 0.08
(m) [UI]
σacc(m) [UI]
0.03 0.06
acc
σ
0.02 0.04
0.01 0.02
0 0
0 50 100 150 200 250 0 50 100 150 200 250
m m
0.02 5
σacc(m) [UI]
σacc(m) [UI]
4
0.015
3
0.01
2
0.005
1
0 0
0 50 100 150 200 250 0 50 100 150 200 250
m m
F IGURE 8.5.: RMS values of accumulated σacc (m), absolute σabs (dotted line) and long term
jitter σlt (dashed line). The curves are constructed using a sample size of N =105 .
Configuration σabs [UI] σlt [UI] σper [UI] σmax [UI] σabs,psd [UI]
Default 0.0258 0.0365 0.00440 0.0512 0.0315
Icp =10 µA 0.0513 0.0726 0.00874 0.1020 0.0611
R0 =400Ω 0.0135 0.0191 0.00273 0.0265 0.0170
Pat=11110000 . . . 0.0046 0.0065 0.00089 0.0073 0.0058
TABLE 8.2.: RMS jitter values from figure 8.5. 1 UI = 333 ps.
figure 8.4, together with calculated values for σabs (dotted line) and σlt (dashed line). In the course
of σacc (m) especially a periodicity given by the spectral peak can be noticed.
Table 8.2 lists the RMS values from equation (8.15). The absolute jitter can also be determined
from the measured phase noise spectrum using Parseval’s Theorem [43, Annex D], and yields
σabs,psd in the last column. Measured RMS values are generally larger than simulated ones, which
is due to mismatches in the peak region and additional parasitic spurs. As can also be observed, an
increase of the charge-pump current leads to a linear increase of absolute jitter and the oscillation
magnitude in figure 8.5.
128
8.2. J ITTER AND P HASE N OISE A NALYSIS
soidal jitter of frequency fSJ and amplitude ASJ at the CPLL input. The BERT-Scope determines
the timing budget of jitter distributions at the output, using a built-in tail fitting method. In order
to get a reference value, the first point is measured at a sinusoidal jitter frequency of typically
fSJ =10kHz, so that the loop filter can easily follow the input jitter and thus, a transfer character-
istic of T (fSJ )=! 1 can be assumed. The corresponding timing budget can subsequently be used as
a reference for determining the complete T (fSJ ) curve over varying frequency fSJ .
In simulations, the same approach requires the use of the tail fitting method from chapter 4. That
is, distributions are gathered from IO jitter values of the behavioral model and extrapolated, in
order to identify the timing budget at the target BER=10−12 . The typical sample size of simulated
distributions is significantly smaller (typically 4 to 5 orders of magnitude) compared to real-time
BERT measurements. This is an essential drawback for simulations, since the timing budget has
to be estimated from higher BER levels, where the number of Gaussian tail samples may not be
sufficient. Further, the fitting algorithm of the BERT-scope may produce different estimates.
Another problem domain can be highlighted with phase noise spectra, when additional random
jitter (RJ) is added to the data input. Figure 8.6 yields a clear amplitude difference between
measured and simulated PSDs. This can be due to model inaccuracies, non-ideal behavior of the
−70
−80
Phase Noise PSD [dBC/Hz]
−90
−100
−110
−120
−130
−140
2 4 6 8 10 12 14
Frequency [Hz] 7
x 10
F IGURE 8.6.: Phase noise PSD mismatch with RJ=0.3 UI and default parameters (table 8.1).
RJ generator, or additional phase noise as caused by interconnect wires and the analog receiver
circuit. Therefore, an additional, alternative analysis method is preferred for model simulations,
in order to truly reflect the transfer characteristic. Such a method is here realized using the FFT
spectrum of absolute jitter at the PLL output. The characteristics of both tail fitting and spectral
methods are briefly summarized:
Tail Fitting Method For each calculated frequency point, N =106 IO-jitter samples are col-
lected and the scaled Q-normalization (sQN) method is applied (ĉ1.2 optimization, ∆Pt =102 ),
while the PLL model is stressed with the jitter frequency fSJ . Only estimated DJpp values are
used for determining the jitter transfer function T (fSJ ). The first frequency value is assumed to
be ideally followed by the loop filter, and thus DJpp (fSJ,min ) is used as a reference to calculate
Spectral Method Transfer function values are calculated from the frequency bins of the jitter
spectrum, which are calculated by the FFT of absolute output jitter. For each frequency point, N
samples are collected and FFT transformed, while the PLL model is stressed with the correspond-
ing sinusoidal jitter frequency fSJ and amplitude ASJ . A sample size of N ≥105 is chosen to
allow for coherent sampling:
NC /N = fSJ /fs (8.17)
129
8. A N ACCURATE B EHAVIORAL M ODEL FOR H IGH -S PEED PLL S
5 5 5
0 0 0
[dB]
[dB]
[dB]
−5 −5 −5
−15 6 7 8
−15 6 7 8
−15 6 7 8
10 10 10 10 10 10 10 10 10
Frequency [Hz] Frequency [Hz] Frequency [Hz]
(a) Default, ASJ =0.4 UI (b) ASJ =0.2 UI (c) ASJ =0.1 UI
measured sQN method FFT method
5 5 5
0 0 0
[dB]
[dB]
[dB]
−5 −5 −5
−15 6 7 8
−15 6 7 8
−15 6 7 8
10 10 10 10 10 10 10 10 10
Frequency [Hz] Frequency [Hz] Frequency [Hz]
(a) Icp =10 µA, R0 =700Ω (b) Icp =20 µA, R0 =700Ω (c) Icp =5 µA, R0 =1600Ω
measured sQN method FFT method
where NC is the bin where the transfer function value is determined. If for example fSJ =1 MHz,
fs =3 GHz ⇒ NC =34, N =102000. Further,
The jitter amplitude represents a peak-to-peak value while the FFT produces a double-side spec-
trum of half amplitude and thus, ASJ must be divided by four.
Both simulation methods use a logarithmic grid fSJ =[105 , 5·107 ] MHz of 50 frequency points.
With the Intel 2.2GHz laptop and N =105 , a full transfer function can be determined in a few
minutes. The default model parameters were already given in table 8.1, with an additional default
jitter amplitude of ASJ =0.4 UI.
An initial comparison of transfer functions over different jitter amplitudes ASJ is given in fig-
ure 8.7, where the typical low pass behavior is observed. Since the CPLL is a non-linear system,
the cut-off frequency of T (fSJ ) varies too. Both simulation methods are able to correctly track
the cut-off frequency. Although the simulated curves generally highlight acceptable matching with
the measured ones, the sQN method generally underestimates the true jitter transfer behavior. This
can especially be observed in the transition region of figures 8.7(c) and 8.8(c). However, the FFT
method correctly reflects the measured peaks. In figures 8.7 and 8.8 generally, both simulation
methods do not exactly match the slope of T (fSJ ) in the cut-off region. It is likely that this effect
ranges from additional parasitic poles which have not been considered, such as the line termina-
tion, input amplifier or equalizer circuit. Figure 8.9 depicts T (fSJ ) for different test patterns, as
130
8.3. S UMMARY
5 5 5
0 0 0
[dB]
[dB]
[dB]
−5 −5 −5
−15 6 7 8
−15 6 7 8
−15 6 7 8
10 10 10 10 10 10 10 10 10
Frequency [Hz] Frequency [Hz] Frequency [Hz]
(a) Default, Pat = 0101 . . . (b) Pat = 0011 . . . (c) Pat = PRBS7
5 5 5
0 0 0
[dB]
[dB]
[dB]
−5 −5 −5
−15 6 7 8
−15 6 7 8
−15 6 7 8
10 10 10 10 10 10 10 10 10
Frequency [Hz] Frequency [Hz] Frequency [Hz]
(d) Pat = LBP (e) Pat = HTDP long (f) Pat = LTDP long
measured sQN method FFT method
specified in [118, section 7.2.4]. Unlike the FFT method, the sQN tail fitting method is now able
to correctly follow the measured course, due to the same analysis principle.
8.3. Summary
A fast system level model for accurate behavioral simulations of charge-pump PLLs has been
developed. Unlike modeling approaches with uniform time steps, an event-driven model uses
state variables that reflect the real physical state and thus, allow for dynamic run-time variations.
The model forms an enhanced version of a prior approach [50] with included gain regulator pole,
VCO noise model and non-ideal behavior of the BB-PD. Depending on the transition density of
the selected data pattern, it is typically able to collect 10-20k jitter values per second simulation
time on an Intel Core Duo 2.2GHz laptop.
Simulation results of the closed loop phase noise over varying parameter configuration proved
excellent agreement with jitter measurements from an existing hardware structure (figure 8.4).
In addition to the PSD spectra, also the RMS values of accumulated jitter were determined in
figure 8.5. They especially highlight the observed PSD peaks.
Two simulation methods were compared with measured jitter transfer functions T (fSJ ) in fig-
ures 8.7-8.9. One of the two simulation methods is based on the equivalent measurement principle,
where a tail fitting method identifies the DJ peak-to-peak characteristics DJpp over varying jitter
frequency. Using a reference value at low frequencies, equation (8.16) allows for determining
T (fSJ ). In measurements, DJpp was determined with a BERT scope, while for simulations the
sQN method from chapter 4 was utilized. As it can be difficult to observe certain phase noise
effects with the sQN method in simulations, a second analysis method based on spectral analy-
sis was implemented as well. It calculates the transfer function by identifying the amplitude of
absolute jitter at the PLL output.
131
8. A N ACCURATE B EHAVIORAL M ODEL FOR H IGH -S PEED PLL S
Both methods correctly identify the cut-off frequency of measured T (fSJ ) in figures 8.7-8.9.
The spectral method is further able to track the peaking behavior of jitter transfer functions, which
is generally underestimated by the sQN method. Transfer characteristics of different test pattern
are instead only reflected correctly by the sQN method, due to the same analysis principle.
The presented model and simulation results have also been published in [C2].
132
9. A Method for Fast Jitter Tolerance
Analysis
Jitter analysis methods can be used for identifying jitter tolerance (JTOL) curves of high-speed
PLLs. This chapter focuses on this application field and realizes an algorithm for the automatic
determination of such curves. Due to the influence of timing jitter along the transmission channel,
high-speed PLLs and CDR structures have to provide a certain robustness against timing varia-
tions. Thus, interface standards often specify tolerance masks [43] to define minimum bounds for
jitter tolerance which must be guaranteed by a system design.
A JTOL curve describes the robustness of a PLL against an injected sinusoidal jitter frequency.
Therefore, a jitter amplitude must be determined where the TJ budget exactly covers the complete
bit period or UI. This corresponds to an inverse problem which is additionally influenced by the
statistical variation of collected distributions. Typically, a JTOL measurement scheme as depicted
in figure 9.1 uses a modulated clock source with corresponding pattern generator, and is charac-
terized by the injected sinusoidal jitter frequency fSJ and amplitude ASJ . The CDR under test
suffers from the jittery signal and produces output data with increased error probability. A bit error
rate tester (BERT) may compare the recovered data with the expected original one, and determine
the resulting error rate. Equivalently, a time interval analyzer (TIA) can directly measure the time
difference between the zero crossing of the analog input signal and the recovered clock edge. The
obtained IO jitter values (section 2.1.2) again allow for collecting distributions that represent the
error rate.
Pattern Jitter
Generator PLL TIA
Recovered
Clock
Delay
Clock
Source Latch
Q BER
BERT
fSJ ASJ D Recovered
Modulation Data
frequency CDR Under Test
The BER measurement principle has already been discussed together with built-in jitter mea-
surement (BIJM) systems in chapter 5, which also explains the similarity between figures 9.1
and 5.1. Again, the adjustable delay element introduces a time discretization which yields a lim-
ited number of bins R per UI. The BERT measurement is easy to implement since bit errors can
be counted. The result is a single probability value, meaning that a complete jitter distribution
requires a sequential BER scan over all R delay steps. A TIA measurement instead, collects jitter
values at every bit transition of the received data and thus, directly yields probability distributions
with maximum speed. It can also be realized using R BERT elements in parallel, which requires
a delay line together with a bank of latches and counters. Although area consuming, BIJMs with
133
9. A M ETHOD FOR FAST J ITTER TOLERANCE A NALYSIS
such a real-time TIA feature have already been realized successfully using high resolution time-
to-digital converters [16, 66].
Jitter tolerance measurements are very challenging, because the amplitude of sinusoidal jitter
must exactly produce the target BER at the CDR under test. Thus, one is searching for the ASJ
value where both distribution tails cross each other at the 10−12 level. So far, this JTOL test
problem has been addressed either from a measurement or a simulation perspective. Methods for
hardware measurements have been proposed in [32, 33, 140, 141], where especially the principle
in [33] is very efficient as it is based on the Gaussian Q-normalization with subsequent tail extra-
polation. Using several measurement points, a regression line can be constructed where the correct
ASJ value is estimated easily via extrapolation. This BERT based approach does not require the
delay element in figure 9.1, but needs a considerable amount of measured bit errors, and is thus
too time consuming for simulations if the whole jitter tolerance curve has to be identified over
varying jitter frequency fSJ . Simulation methods work equivalent to TIA based measurements.
They additionally use statistical models [97] or special waveforms [4] to minimize the required
number of jitter samples as much as possible, so that JTOL simulations can be carried out in a
feasible amount of time. However, they can generally not be used for hardware measurements.
In this chapter an analysis method is proposed, where the jitter tolerance curve of a PLL is
determined very quickly using an adaptive algorithm. The method is sufficiently fast for use with
both behavioral simulations and TIA based jitter measurements. A minimum measurement time
is obtained by automatically adapting the sample size of collected jitter distributions according to
the dynamics of the PLL under test. This adaptive recursion utilizes the previously described sQN
and QN methods from chapters 4 and 6.
In subsequent sections, first the adaptive principle of the algorithm is described. In a practi-
cal use case, then a simulation example reuses the modeled charge-pump PLL from chapter 8.
Obtained simulation results show, that JTOL curves can also be determined correctly when using
hardware systems of limited accuracy. In a final analysis, simulation results are compared with
JTOL measurements from the given hardware structure.
This equation is widely used in adaptive filter theory [49] where least-mean-square algorithms or
134
9.1. A DAPTIVE A LGORITHM FOR JTOL A NALYSIS
Kalman filters are implemented, and offers a high robustness against statistical variations of the
error term e(n). Using a block size of N jitter values, a tail fitting algorithm can provide a cost
function and is thus able to specify the error e(n). The jitter amplitude for the next iteration step
ASJ (n + 1) is then determined according to equation (9.1), using the old value ASJ (n) and the
error e(n) which is additionally scaled by the learning rate parameter ν.
The second part of the algorithm adaptively adjusts the sample size N for each collected jitter
distribution. The method starts with a minimum sample size Nmin and decides after each itera-
tion, whether N must be increased or not. As soon as a maximum number Nmax is successfully
reached, the search algorithm converges and the JTOL analysis can proceed with the next jitter
frequency. This second algorithm allows for a significant speed-up of the ASJ search, because
only few jitter samples are needed for initial iterations. Further, the increase of N starts at a point
where ASJ is already quite close to the final result.
For the QN method Q(p)=Qest is directly calculated from the left hand side, while for the sQN
method it can be determined recursively using a simple Newton iteration. Qest must approach the
desired target BER=10−12 , which gives the normalized error term e(n):
This result is used together with the adaptive algorithm in equation (9.1).
135
9. A M ETHOD FOR FAST J ITTER TOLERANCE A NALYSIS
vSJ [L]=ASJ (n + 1)
L=L+1
n o
s
{v,min , Lmin } = min v = t(a, L−1) · √ v,L
L·vL
vSJ = vSJ [L−Lmin , . . . , L−1]
no fp (N ) yes
v,min < conf · End
fp (Nmax )
no yes
v,min
N =fp−1 conf
· fp (Nmax ) N =Nmax ASJ = vL
block of jitter samples has been collected. After applying the tail fitting method, the error e(n)
from equation (9.4) is determined and used for updating the recursion in equation (9.1). The re-
sulting new ASJ (n + 1) value is stored in an array vSJ [0, . . . , L−1] of variable length, where the
statistical variation of amplitudes can be observed over multiple iterations.
With blocks of only Nmin jitter samples at the beginning, the recursion quickly settles ASJ (n)
to a level where it constantly oscillates around its true value and exhibits statistical random walks.
A measure for the statistical variation of ASJ (n) can be derived if only the L last recursions
are considered. Assuming a normal distribution, the confidence interval of ASJ is specified as
t-statistic with
sv,L
v = t(a, L −1) · √ (9.5)
L · vL
and
1X
vL = ASJ,i (9.6)
s PL
A2SJ,i − L · vL 2
sv,L = (9.7)
L−1
where v is the estimated confidence bound of a t-distribution with confidence level a=0.95 and
L−1 degrees of freedom. It is proportional to the empirical standard deviation sv,L and normalized
by the empirical mean vL . If v falls below the target deviation conf , the JTOL algorithm has
converged.
The length L of the array vSJ is continuously increased. It is incremented at every recursion,
but only a subset of its newest elements is used to minimize the observed statistical variation. At
each recursion, the minimum epsilon value is searched over all possible lengths:
This minimum search yields an optimistic estimate of the actual statistical confidence of ASJ
values. It allows for quickly changing to a higher sample size N as soon as the observed optimistic
tolerance v,min falls below a known comparison threshold. Hence, the algorithm behavior is
optimized with respect to a minimum number of recursions.
136
9.1. A DAPTIVE A LGORITHM FOR JTOL A NALYSIS
0.1 0.03
0.08
0.02
0.06
Emed
σe
QN QN
0.04
0.01
0.02
sQN sQN
0 4 5 6 7 8
0 4 5 6 7 8
10 10 10 10 10 10 10 10 10 10
N N
F IGURE 9.3.: Worst case error behavior of tail fitting methods over varying N . 4th order regres-
sion polynomials, K=250, worst case distribution shapes ADJ /σRJ =1/2 (sQN)
and 1/4 (QN).
Alg. p0 p1 p2 p3 p4
QN 0.2036 −0.03269 0.001823 −3.466·10−5 0.0
sQN 0.3493 −0.08615 0.008218 −3.530·10−4 5.71·10−5
The ideal comparison threshold corresponds to the expected error of the tail fitting method.
This error behavior can only be approximated, because extrapolation results depend not only on
the sample size N , but also on the underlying distribution shape. The CDR under test is stimulated
with sinusoidal jitter and thus, collected jitter distributions are expected to consist of a bounded
sinusoidal component combined with Gaussian random jitter. From the performance comparison
in figures 6.11-6.15, worst case distribution shapes resulting from combined sinusoidal and Gaus-
sian jitter components can be specified for the two fitting methods. The worst case error for each
method can thus be used to obtain a simplified function of sample size N . In figure 9.3 the median
error Emed (left) and corresponding standard deviation σe (right) curves are plotted. K=250 tail
fits were carried out for each of the different sample sizes, and 4th order polynomials were fitted to
achieve a functional relationship fp (N ) between sample size N and error behavior. Note, that the
error bias Emed cannot be compensated with the JTOL algorithm since the underlying distribution
shape is basically unknown. However, the standard deviation σe can be used as a pessimistic in-
dicator for choosing the right sample size N . The logarithmic scaling in figure 9.3(b) leads to a
polynomial
fp (N ) = p0 + p1 · log(N ) + . . . + p4 · log(N )4 (9.9)
with the coefficients in table 9.1, and an analysis range of N =[104 , 108 ]. In order to compare the
actual confidence interval v,min with the expected error of the fitting method, we can use these
error polynomials and formulate the condition
fp (N )
v,min < conf · (9.10)
fp (Nmax )
The error at fp (N ) is normalized by fp (Nmax ) so that the target bound conf forms the refer-
ence. In the flow graph (figure 9.2) this condition decides whether the obtained ASJ estimates
are sufficiently accurate, so that jitter distributions of a larger sample size have to be collected in
subsequent iterations. If this is the case, a new value for N is determined from the inverse of the
137
9. A M ETHOD FOR FAST J ITTER TOLERANCE A NALYSIS
actual v,min , otherwise the JTOL algorithm continues to iterate. The inverse fp−1 (N ) is simply
realized by a Newton approach.
The overall structure of the algorithm guarantees for a strictly monotonic increase of N until
Nmax is reached. With v,min <conf and N =Nmax the final convergence criterion is met. Note,
that in order to identify a complete jitter tolerance curve, the JTOL algorithm from figure 9.2 must
be repeated for every desired frequency fSJ . Some additional speed up can thus be achieved when
using the amplitude result of the last frequency as initial value for the next one.
An additional reset feature has also been included with the presented JTOL algorithm. Since the
jitter amplitude ASJ is operated in a region where the observed jitter extends over the complete
UI, the modeled PLL may easily become unstable and produce bit errors. Thus, the JTOL method
also requires a well defined reset behavior. This is realized by first returning to a smaller, stable
ASJ value. Then the learning rate parameter ν is additionally decreased by a factor of two, so that
the search algorithm is more focused on the region of interest.
138
9.2. A PPLICATION E XAMPLE
6
0.8 10
0
3 10
0.6
2 −1
10
εv,min
5
en
ASJ
N
0.4 10
1 −2
10
0.2
0 −3
10
4
0 10
0 5 10 15 20 0 5 10 15 20 0 5 10 15 20 0 5 10 15 20
Iteration n Iteration n Iteration n Iteration n
(a) ASJ (n), fSJ =100MHz (b) e(n), fSJ =100MHz (c) v,min (n), fSJ =100MHz (d) N (n), fSJ =100MHz
6
0.8 10
0
3 10
0.6
2 −1
10
εv,min
5
en
ASJ
N
0.4 10
1 −2
10
0.2
0 −3
10
4
0 10
0 5 10 15 20 0 5 10 15 20 0 5 10 15 20 0 5 10 15 20
Iteration n Iteration n Iteration n Iteration n
(e) ASJ (n), fSJ =10MHz (f) e(n), fSJ =10MHz (g) v,min (n), fSJ =10MHz (h) N (n), fSJ =10MHz
4 6
10
0
3 10
2 −1
10
ASJ
εv,min
5
en
N
10
1 −2
10
0 −3
10
4
0 10
0 5 10 15 20 0 5 10 15 20 0 5 10 15 20 0 5 10 15 20
Iteration n Iteration n Iteration n Iteration n
(i) ASJ (n), fSJ =2MHz (j) e(n), fSJ =2MHz (k) v,min (n), fSJ =2MHz (l) N (n), fSJ =2MHz
sQN QN
F IGURE 9.4.: Examples for the adaptive JTOL algorithm converging toward the unknown jitter
amplitude ASJ . Parameters are given in table 9.2.
will require more iterations before the obtained statistical confidence is sufficiently accurate. In
extreme cases, it may even become unstable and produce errors. Otherwise, if ν is chosen too
small the JTOL search will converge before ASJ reaches the correct value, simply because also
the amplitude variations are too small. The ideal ν depends on the dynamics of the investigated
system and thus, changes with varying jitter frequency fSJ as well as the parameter configuration
of the model.
In figure 9.5 the convergence probability of the JTOL algorithm is investigated over varying ν.
As model configurations, four different parameter settings as given in table 9.3 were used, each
with twenty frequency points in a logarithmically scaled range of fSJ =[106 , 108 ] MHz. Success-
ful JTOL runs were counted by combining two limiting criteria: number of iterations I and overall
sample size N . In order to yield a valid JTOL run, both criteria must be fulfilled.
The obtained curves in figure 9.5 thus represent an empirical probability for convergence. They
clearly highlight an increase in the 0.01≤ν≤0.1 region where the performance settles toward a
maximum. When further increasing ν, a larger variation of probabilities is observed because the
obtained ASJ values also suffer from a strong statistical fluctuation. At largest ν, the PLL further
looses its lock state several times. Thus the reset feature, as described previously, divides ν until a
stable convergence of ASJ values is again achieved. For the presented PLL model the learning rate
is selected such that maximum performance with the QN method is achieved, which is at ν=0.11.
Further, a maximum number of iterations Imax =50 aborts simulations if no convergence criterion
is reached. All default settings for the JTOL algorithm are again summarized in table 9.2.
139
9. A M ETHOD FOR FAST J ITTER TOLERANCE A NALYSIS
0.8
P (I ≤ 30 ∧ N ≤ 20·106 )
P (I ≤ 25 ∧ N ≤ 15·106 ) sQN
Probability
0.6 P (I ≤ 20 ∧ N ≤ 10·106 )
0
0 0.05 0.1 0.15 0.2
ν
140
9.2. A PPLICATION E XAMPLE
R = 32 R = 512
Param. Alg. Icp R0 Kv Icp R0 Kv
Def. Def.
10µA 400Ω 4GHz/V 10µA 400Ω 4GHz/V
QN 89.2 102.3 146.3 86.1 84.4 93.3 84.8 84.9
sQN 88.9 105.6 94.9 80.2 67.0 65.0 85.2 65.2
N [M]
c-QN 182.0 294.0 229.0 269.0 175.0 285.0 209.0 256.0
c-sQN 193.0 301.0 211.0 247.0 187.0 284.0 221.0 240.0
QN < 0.04 < 0.01 < 0.03 < 0.01 0.12 0.10 0.18 0.13
sQN 0.32 0.39 0.39 0.28 2.30 3.11 2.73 2.71
tc [s]
c-QN < 0.02 < 0.03 < 0.04 < 0.02 < 0.05 0.12 < 0.09 0.11
c-sQN 0.31 0.37 0.33 0.30 1.83 2.79 2.27 2.37
QN 13.8 21.2 17.7 16.6 13.7 20.5 15.6 17.2
sQN 15.3 19.2 17.7 16.3 13.9 18.8 15.5 16.9
If
c-QN 9.1 14.7 11.5 13.5 8.8 14.3 10.5 12.8
c-sQN 9.7 15.1 10.6 12.4 9.4 14.2 11.1 12.0
sQN are slightly better than those of QN, due to the faster convergence. Both c-QN and c-sQN
methods with constant N yield the best results for If , because also the error bias is constant (refer
to figure 9.3(a)).
Table 9.4 shows the behavior of the JTOL algorithm, when carried out with a coarse time reso-
lution or reduced number of bins R, as is the case for hardware measurements. Generally a slight
increase of the number of overall samples N is observed, especially with sQN and QN methods.
This is, because a coarse time resolution R leads to larger statistical variations of TJ estimates,
obtained from extrapolated distribution tails. Thus, the JTOL algorithm on average also needs
more samples in order to reach the convergence criterion.
Overall sample size N and average number of iterations If of both QN and sQN methods can
also be investigated over varying R, as shown in figure 9.6. While If remains approximately con-
stant over the complete analysis range, N in fact highlights a slight increase toward smaller R.
This empirically confirms the assumption of a larger statistical variation of TJ estimates. How-
ever, at any R the JTOL algorithm correctly converges within the specified maximum number of
iterations Imax =50, indicating that it can also be utilized for hardware measurements.
141
9. A M ETHOD FOR FAST J ITTER TOLERANCE A NALYSIS
20
16
14
12
7
10 1 10 1 2 3 4 5
2 3 4 5
10 10 10 10 10 10 10 10 10 10
Number of bins R Number of bins R
sQN QN
F IGURE 9.6.: a) Average number of iterations If and b) total sample size N over varying R.
1 1 1 1
10 10 10 10
SJ
ASJ
ASJ
ASJ
0 0 0 0
10 10 10 10
A
−1 −1 −1 −1
10 6 7 8
10 6 7 8
10 6 7 8
10 6 7 8
10 10 10 10 10 10 10 10 10 10 10 10
fSJ fSJ fSJ fSJ
(a) sQN, Default (b) sQN, Icp =10 µA (c) sQN, R0 =400Ω (d) sQN, Kv =4GHz/V
1 1 1 1
10 10 10 10
SJ
ASJ
ASJ
ASJ
0 0 0 0
10 10 10 10
A
−1 −1 −1 −1
10 6 7 8
10 6 7 8
10 6 7 8
10 6 7 8
10 10 10 10 10 10 10 10 10 10 10 10
fSJ fSJ fSJ fSJ
(e) QN, Default (f) QN, Icp =10 µA (g) QN, R0 =400Ω (h) QN, Kv =4GHz/V
measured R=Rsim =3.3 · 105 R=128
F IGURE 9.7.: Simulated (tables 9.3 and 9.4) and measured JTOL curves at different model pa-
rameter configurations.
With the confidence level conf =0.005, this is further achieved without visibly affecting the
quality of obtained JTOL curves. In figure 9.7 the simulated JTOL curves for the sQN and QN
methods at both Rsim and R=128 are plotted together with manually measured JTOL curves of
the same PLL hardware structure. Note that the measurement device (SyntheSys BERT-Scope
7500A) has an amplitude limit of ASJ,max = 3.3 UI, which is reached at fSJ =1 MHz. Further,
differences with simulated curves are mainly given by model inaccuracies and thus, do not re-
flect the performance of the JTOL algorithm. The hardware oriented model simulation with only
R=128 bins already matches excellent with the Rsim high resolution scenario and can still handle
the varying loop dynamics over the complete frequency range. The obtained results show, that
the best jitter tolerance curve is achieved with the default model parameter settings. The S-ATA
mask specification [118] demands a minimum tolerance of ASJ = 0.42 UI which is here clearly
guaranteed.
If a complete JTOL measurement over twenty frequency points with the QN method typically
requires a total of N ≈ 100M samples (≈ 1.2h of simulation time), the time consumed in a TIA
based hardware measurement is tN ≈ 33ms for a 3Gb/s interface. Together with tc ≈ 130ms
(R = 512) for calculations and 1ms additional time buffer per iteration tI ≈ If · 20 · 1ms = 320ms
142
9.3. S UMMARY
which is needed to identify the complete jitter tolerance curve over twenty frequency points.
9.3. Summary
A fast and accurate method for the identification of jitter tolerance curves of high-speed PLLs has
been presented. An adaptive algorithm determines the unknown jitter amplitudes recursively and is
optimized with respect to a small number of iterations and sample size. The algorithm realization
started with the simple adaptive recursion in equation (9.1) and the associated derivation of a cost
function given in equation (9.4). An algorithm for automatic sample size adaptation was realized
and described with the flow graph in figure 9.2.
The basic idea is to observe jitter amplitudes over a subset of L last recursions, which allows
to describe their statistical confidence interval in terms of a Gaussian t-statistic (equation (9.5)).
As soon as this value falls below a predefined threshold, the algorithm adapts the sample size N
of collected distributions accordingly. This adaptation process is controlled by the known error
behavior of QN and sQN fitting methods. Therefore, the worst case distribution shapes of both
methods were used to derive 4th order polynomials, which approximate the extrapolation error as
a function of sample size N (figure 9.3). The overall algorithm structure guarantees for a strictly
monotonic increase of the adapted sample size until the maximum Nmax is reached, together with
the desired confidence level. The presented algorithm is repeated for every jitter frequency fSJ .
As an application example, the proposed analysis method was applied to the PLL model from
chapter 8. First, suitable analysis parameters were identified to optimize the algorithm behavior
with respect to the given model. As an example, figure 9.4 highlights the varying loop dynamics of
the PLL at different fSJ . Default algorithm settings, such as the optimized learning rate parameter
ν=0.11 were specified in table 9.2.
Finally, the analysis proceeded to the determination of jitter tolerance curves. Therefore, four
algorithmic versions were compared, including the two fitting methods with (QN and sQN) and
without (c-QN and c-sQN) sample size adaptation. Performance results were given in table 9.3 for
simulators (R=Rsim =3.3·105 ), and in table 9.4 for simulated hardware measurements (R=512
and R=32). These include four different parameter settings with twenty fSJ values each.
Results demonstrated, that the adaptive sample size adaptation of QN and sQN generally de-
creases the overall number of required jitter samples N by a factor of 2-3. This value directly
reflects the simulation effort or the measurement time in a hardware system. The overall computa-
tion time tc is the smallest with QN and c-QN methods and outperforms sQN and c-sQN methods
for more than one order of magnitude. Finally, the average number of recursions per jitter fre-
quency If is the smallest without sample size adaptation (c-QN and c-sQN methods). This is due
to the constant error bias which does not influence the adaptation process, as is the case with QN
and sQN methods. The same performance characteristics of N , tc , and If can also be observed
with simulated hardware scenarios, besides a highly reduced computational demand. For each
test case in figure 9.6, the algorithm was always able to converge within a maximum number of
fifty recursions over varying R, thus indicating that the proposed algorithm can also be applied to
hardware measurements.
As a final result, in figure 9.7 the obtained jitter tolerance curves were compared against real
jitter measurements from the test structure and showed an excellent matching. The simulation
of such jitter tolerance curves is particularly useful to optimize a CPLL design with respect to
its robustness against input jitter. It also allows for the verification of imposed specification re-
143
9. A M ETHOD FOR FAST J ITTER TOLERANCE A NALYSIS
quirements such as a jitter tolerance mask. Contents of this chapter have partly been published in
[C6].
144
10. An FPGA based Diagnostic Tool for
Jitter Measurement and Optimization
In order to highlight the practical aspects of tail fitting methods, in this chapter an embedded jitter
measurement system is presented, which acts as a diagnostic tool for serial high-speed interfaces.
The underlying idea is to combine a real BIJM system with the previously described sQN and
QN tail fitting methods. This is to confirm the theory of hardware design aspects from chapter 5
and to prove correctness of the derived equations and empirical relations. Further, extrapolated
tails yield the TJ timing budget, which allows for judging the quality of transmission lines, PLLs
or transceiver structures as system under test (SUT). The resulting diagnostic tool is thus able to
optimize and configure an SUT without the use of additional instrumentation devices.
Many BIJM topologies and embedded systems [16, 18, 57, 64, 65, 68, 73,79] have been designed
for production tests and on-chip diagnostics, and some of them have also been used for jitter
optimizations [87, 132]. However, the combined use of BIJMs with tail fitting methods has not
been considered so far. This chapter especially points out the benefit of applied tail fitting methods
to estimate the TJ budget, which forms a direct quality measure for the impact of timing jitter on
system performance. The target platform is a Virtex-5 FPGA [137] on an ML507 high-speed
evaluation board [138]. It produces a 3Gb/s serial reference signal and retrieves timing jitter
information from the SUT.
In the following sections first the diagnostic principle is introduced. Then the implemented
FPGA logic together with the analysis software is described, and parameter optimizations with
different diagnostic scenarios and test cases are carried out. Finally the observed tail fitting error
is compared with expected worst case errors from the theory in chapter 5. A brief summary is
given at the end.
145
10. A N FPGA BASED D IAGNOSTIC TOOL FOR J ITTER M EASUREMENT AND O PTIMIZATION
Tx
Data
Ref Clk
System
Under Delay
Test
(SUT)
PD Counter
Jitter
BIJM Analysis
measurement runs of varying delay. The smallest delay step defines the number of bins R in a UI,
while the PD can also be replaced by a simple D flip-flop if the input data stream is known. In this
case the recovered data is first compared against the expected transmit data, which then allows for
error counting.
If a jitter distribution is measured down to a BER level of 1/N over R delay steps, the overall
time consumed is
tt [UI] = N · R. (10.1)
Although test time is rather uncritical, also in diagnostic applications the measured BER depth
1/N poses a fundamental time limit. In order to estimate the TJ budget at the target BER=10−12 ,
again tail fitting methods have to be applied. Using the sQN and QN methods from chapters 4
and 6 allows to parameterize the Gaussian tails (µ, σ, A) of a jitter distribution, and thus, to
estimate the extrapolated TJ peak-to-peak value TJpp
TJpp = tR − tL (10.2)
−12
tL(R) = µL(R) − σL(R) · Q(10 /AL(R) ) (10.3)
with the Q-function defined in equation (3.10). The obtained TJpp estimates are subsequently
used for jitter diagnostics and parameter optimization.
10.1.2. Implementation
Subsequently a 3Gb/s jitter diagnosis system is implemented using an FPGA. The measurement
is realized using a Xilinx Virtex-5 FX70T on an ML507 high-speed evaluation board [138], which
is controlled by a MATLAB program running on a remote desktop computer. The general FPGA
block scheme is given in figure 10.2. In order to collect jitter distributions, a dedicated high-speed
transceiver (GTX) [137] is combined with a BER test (BERT) logic and controlled by an instanti-
ated microprocessor (MP). Measurement results are then transferred to the remote computer where
the tail fitting methods are applied. The three main blocks of the FPGA logic are subsequently
described in more detail.
High-Speed Transceiver
The Xilinx GTX is used for the parallel to serial and serial to parallel data conversion, and is
needed to lower the data rate for use with the internal, custom FPGA logic. The 3Gb/s data stream
is converted from or into a 20bit word along with a 150MHz clock signal to mark the beginning
of a new word. The transceiver is driven in lock-to-reference mode, meaning that the PLL for data
recovery (Rx-PLL) is locked to the same reference frequency of the transmitter buffer (Tx-PLL),
146
10.1. M EASUREMENT P RINCIPLE AND I MPLEMENTATION
F IGURE 10.2.: Block scheme for the FPGA based 3Gb/s jitter measurement system.
in order to realize the diagnostic principle from figure 10.1. A complete jitter distribution can be
measured by adjusting the data sampling point of the receiver PLL over R=128 time steps. This
is done by using the Dynamic Reconfiguration Port (DRP) [137] of the GTX, which is controlled
by the MP. The transmit signal is also a 20bit word given by the MP.
BER Tester
The custom made BERT logic consists of two units: a bit counter and an error counter (figure 10.3).
The former keeps track of the sum of total bits by counting the number of clock cycles from the
Rx clock. The latter compares the Rx data to the data pattern set by the MP. The number of
errors is then added every Rx clock cycle. Each of the two counters is compared to a preset
maximum value. Once either number is reached, a generated enable signal simultaneously stops
both counters. This way the measurement can be stopped by either the number of bits or errors,
which is especially advantageous during the synchronization phase (MP). Once a measurement at
one sampling position is completed the MP reads the values held by the counters, resets them and
increases the delay to start the next measurement run.
Microprocessor
The MicroBlaze Processor controls the measurement sequence and realizes the serial interface
(RS232) between the ML507 board and a remote computer. The software flow graph in figure 10.4
shows the steps of a measurement run, needed for collecting a complete jitter distribution. The
data pattern to be transmitted along with the number of bits N per sampling position are first
>=
Rx
20x XOR 16bit Error Comp
Accumulator
max.
RST Bits OR
Rx−Clk EN
>=
Comp
36bit
BERT logic Counter RST
EN INV
147
10. A N FPGA BASED D IAGNOSTIC TOOL FOR J ITTER M EASUREMENT AND O PTIMIZATION
Synchronization
Error
Measurement
Samp.Pos.
stops if either:
+1
− max. # errors
− max. # bits
no
128
Data Transfer sampling Apply Tail fitting
to Matlab positions Algorithm
yes
entered using MATLAB. The MP then starts with short BER measurements at the first delay step
in synchronization mode. This phase is needed to correctly match the 20bit Rx and Rx-Cmp
words, due to an unknown channel delay. During the synchronization, the Rx word is compared
with a circular shifted version of the original Tx data pattern, where the BERT collects only 3200
words or 64k bits to evaluate the number of errors. This Rx-Cmp pattern is successively rotated
over all 20 positions. The lowest number of errors thus indicates the synchronization pattern to
which the Rx data pattern needs to be compared. The synchronization phase is carried out very
quickly, consuming < 0.5ms, and is repeated at each of the R sampling positions or delays. As a
key advantage, this allows for measuring both distribution tails in a single measurement run (also
see figure 10.5).
After synchronization, the MP resets both counters, and starts the long BER measurement with
sample size N and a maximum number of 32k errors. Once either counter reaches the maximum,
the MP reads their values and calculates the BER, passes it on to MATLAB, and increases the
sampling delay to restart the measurement process. Once the BER at every delay step has been
measured, MATLAB runs the program containing the sQN and QN fitting methods, which finally
yields the desired TJpp value at the target BER=10−12 .
148
10.2. J ITTER M EASUREMENTS AND O PTIMIZATION
0
10 0
10
N = 108
Pat = 08CEFHex
BER
−4 −2
10 10
−4
10
−8
10
1 32 64 96 128 CDF 1 − CDF
BER
−6
10
20
−8
15 10
Sync Pos
10 −10
10
5
−12
0 10
1 32 64 96 128 −0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15 0.2
Sampling Position Jitter Extent [UI]
(a) (b)
F IGURE 10.5.: Example for a measured jitter distribution (left) and sQN fitted tails (right). N =
108 , Pat = 08CEFhex , 1m RG-58 coaxial cable.
0.45
0.28
RG−58
1m RG−58 8
N = 10
0.27 Pat = 08CEFhex 0.4
Pat = 08CEFhex
0.26 0.35
TJpp
TJpp [UI]
0.25
0.3
0.24
0.25
0.23
0.2 0 1
0.22 10 10
1M 10M 100M 1G 100G Cable length [m]
(a) (b)
F IGURE 10.6.: K=100 evaluations of TJpp estimates over a) varying sample size N , and b)
RG-58 cable length. Triangles mark medians of exact measurements (N = 1011 ).
as can be seen from the example in figure 10.6(a) where K=100 distributions of the 1m RG-58
cable are fitted over varying sample size N . As expected, both bias and statistical spread decrease
with larger N because the extrapolation range becomes smaller. Note that the error bias is always
positive and thus, yields pessimistic TJpp estimates, which is due to the beneficial extrapolation
property described in section 4.2.1.
In figure 10.6(b) K=100 estimates of TJpp are statistically evaluated over varying cable length,
to estimate the influence of ISI on total jitter. The triangles mark median values of N =1011
measurements to approximate the ideal extrapolation result. With 5m and 7m cables, the signal is
additionally degraded by noise effects, which cause a larger error bias in figure 10.6(b).
With N =108 samples a measurement run takes approximately 2s, which is sufficiently fast for
jitter diagnosis. This value is clearly below the expected test time of
This is, because bins with high error rates quickly count bit errors and thus, the maximum number
is reached very fast. Additionally, the FPGA software first locates a jitter distribution within the
128 sampling positions, so that measurements are only made in the region of interest.
The reference clock of the GTX transceiver includes a certain amount of inherent jitter, which
forms the fundamental minimum for the jitter measurement system. It can be identified using
the internal loopback mode of the GTX, and yields typical fitting results for right bathtub tails as
shown in figure 10.7. Due to the clean data signal, tails become very steep and thus, contain only
few data points. Especially at N ={106 , 107 } a correct tail extrapolation is therefore not possible
149
10. A N FPGA BASED D IAGNOSTIC TOOL FOR J ITTER M EASUREMENT AND O PTIMIZATION
0.05 0
QN N = 108
1
Pat = 1010...
0.045
2
Q(BER)
0.04
4
5
0.035
6
7
0.03 0 0.01 0.02 0.03 0.04
1M 10M 100M 1G 10G 100G Jitter Extent [UI]
F IGURE 10.7.: Estimated positive jitter at internal loopback mode, using QN as fitting method.
Especially for smallest sample sizes N ={106 , 107 } a correct tail extrapolation
fails because bathtub curves (right) become too steep.
anymore. At N =108 further, the median is smaller than the approximated true median at N =1011 ,
which is due to the error oscillations at lowest bin density, as described in section 5.4.2. Observing
multiple measurement runs, the linear course in Q-domain is given until Qup =Q(∆Pt /N )≈1,
while σt,10−11 ≈3.7·10−3 . These values can be inserted into equation (5.20) for calculating the
required minimum sample size, which gives Nmin >107 . This is also confirmed by the results in
figure 10.7. The timing budget of GTX inherent jitter is approximated with
Instead of testing the quality of transmission channels, the jitter diagnosis system can also opti-
mize the parameter configuration of a complete high-speed interface as SUT. To demonstrate this
concept, in an example the GTX internal Rx and Tx structures are optimized together with a 5m
RG-58 coaxial cable.
First, only the Tx buffer configuration is optimized. Therefore, the GTX provides three ports
which control differential output swing (TxDIFF), pre-driver swing (TxBUF) and pre-emphasis
(TxPRE) [137]. Each of these ports has eight different gain settings, thus with a total of 512 param-
eter combinations. Using the diagnostic tool, the TJpp value acts as fitness measure for parameter
optimization, and yields the surfaces in figure 10.8. The selected sample size for measurements
is N =108 . Although median values from a statistical evaluation with K=25 measurements are
displayed, a single evaluation is already sufficient for optimizations, since the obtained standard
deviation of estimates is always σT Jpp,est <0.008 UI.
Second, the Rx-PLL is used in normal lock-to-data mode as is the common case for high-speed
data recovery. Without the clean reference clock, the Rx-PLL now suffers from a significant
0.48 TxPRE = 0
0.48 TxPRE = 6
0.44 0.44
TJpp [UI]
TJpp [UI]
TxPRE = 4
0.4 0.4
0.36 0.36
0 0
2 0 2 0
TxPRE = 2 2 2
4 4
4 4
6 6 6 6
TxBUF TxDIFF TxBUF TxDIFF
150
10.3. A NALYSIS OF E XTRAPOLATION E RROR
TxPRE = 4
TxPRE = 0
0.65 0.65
TxPRE = 6
0.6 0.6
TJpp [UI]
TJpp [UI]
0.55 0.55
F IGURE 10.9.: TJpp surfaces for Tx buffer optimization, lock-to-data mode. N =108 , K=25.
0.5 1
Large EQ Medium EQ
0.8 Large EQ
Medium EQ
0.4
Small EQ
TJpp [UI]
TJpp [UI]
Small EQ
0.6
0.3
0.4
EQ Gain Bypass EQ Gain Bypass
0.2 0.2
0 5 10 15 20 25 30 0 5 10 15 20 25 30
DFE1 tap weight DFE1 tap weight
F IGURE 10.10.: TJpp optimization of Rx structures, including four different EQ settings and a
single DFE tap weight. N =108 , K=25.
amount of additional jitter caused by the recovered signal clock, as can also be seen in figure 10.9.
151
10. A N FPGA BASED D IAGNOSTIC TOOL FOR J ITTER M EASUREMENT AND O PTIMIZATION
0.26 0.26
1m S−ATA 1m S−ATA
Pat = 0101... Pat = 0101...
0.25 0.25
0.24 0.24
[UI]
[UI]
pp
pp
TJ
TJ
0.23 0.23
0.22 0.22
0.21 0.21
1M 10M 100M 1G 100G 1M 10M 100M 1G 100G
F IGURE 10.11.: Estimated TJpp values of a 1m S-ATA crossover cable (solid boxes),
N ={106 , . . . , 109 }. The N =1011 measurement allows for worst case error
estimation (equation (5.30), tables of coefficients 5.4 and 6.2) by assuming si-
nusoidal (PLL noise dominated) jitter (dashed boxes). ∆Pt =103 .
0.26 0.26
1m S−ATA 1m S−ATA
Pat = 08CEFhex Pat = 08CEFhex
0.25 0.25
0.24 0.24
TJpp [UI]
TJpp [UI]
0.23 0.23
0.22 0.22
0.21 0.21
1M 10M 100M 1G 100G 1M 10M 100M 1G 100G
F IGURE 10.12.: Estimated TJpp values of a 1m S-ATA crossover cable (solid boxes),
N ={106 , . . . , 109 }. Worst case error estimation with assumed uniform (ISI
dominated) jitter (dashed boxes). ∆Pt =103 .
cable length. With short cables jitter is expected to be dominated by ISI, while for longer cables
additional noise effects such as couplings, reflections or crosstalk will also degrade the signal.
Both QN and sQN fitting methods are first applied to jitter distributions of a standard 1m S-ATA
crossover cable and a clock-like data pattern. A statistical evaluation of TJpp values over K=100
bathtub measurements and varying sample size N yields the solid boxes in figure 10.11. As
expected, with increasing N the obtained medians converge toward the true value while statis-
tical spread becomes smaller. Thus, a good estimate for the true TJ budget is obtained from
measurements at N =1011 (K1011 =25). The observed statistical behavior of TJpp estimates can
be compared with expected worst case errors using the derived empirical relation (5.30). With
σRJ ≈ σt,10−11 ≈ 0.012, these errors can be calculated by choosing the correct DJ type. For
a clock-like pattern, the signal is only affected by the PLL jitter of the reference clock and the
transmission channel cannot produce additional ISI. Thus, sinusoidal DJ is assumed (also refer
to section 3.2.2). The obtained statistical error (dashed boxes) must be generally larger than the
measurements, which is the case and confirms validity of the worst case error estimation. Only
at N =106 , fitted values indicate a larger error bias jump, meaning that the underlying DJ shape
behaves different from the assumed sinusoidal distribution. Thus, the given sample size is not
anymore sufficient for correct tail extrapolation.
152
10.3. A NALYSIS OF E XTRAPOLATION E RROR
0.4 0.4
5m RG−58 5m RG−58
0.39 Pat = 08CEFHEX 0.39 Pat = 08CEF
HEX
0.38 0.38
[UI]
[UI]
0.37 0.37
pp
pp
TJ
TJ
0.36 0.36
0.35 0.35
0.34 0.34
1M 10M 100M 1G 100G 1M 10M 100M 1G 100G
F IGURE 10.13.: Estimated TJpp of a 5m RG-58 coaxial cable (solid boxes), N ={106 , . . . , 109 }.
Worst case errors (dashed boxes) assume quadratic curve DJ (combined ISI and
unknown noise couplings, equation (5.30), tables 5.4 and 6.2). ∆Pt =103 .
The second example in figure 10.12 uses a PRBS4-like 08CEFhex pattern for ISI dominated
jitter, instead of the clock-like signal. The expected worst case error assumes uniform DJ and is
again larger than the measured error.
A third example in figure 10.13 shows the influence of additional noise effects, as appear with
long transmission channels. Therefore, jitter measurements are now carried out with a 5m RG-58
coaxial cable and statistical evaluations are again performed using sQN and QN methods. Also
the PRBS4-like 08CEFhex pattern is used to produce ISI. The worst case error behavior can now
be described as a combined effect of ISI and unknown noise couplings, using quadratic curve
DJ. Validity of this assumption is confirmed with the obtained extrapolation results, which remain
within the expected error limits. At N =109 the worst case error yields somewhat pessimistic
medians and rather tends to underestimate the measured statistical spread. Note that this sample
size is also located beyond the original range of regression planes N =[5·105 , 108 ] for the tables
of coefficients 5.4 and 6.2, which was considered as more suitable for this use case.
The fourth example investigates the influence of process variations on fitted tails. Therefore
the Rx-PLL of the GTX is used in lock-to-data mode which is the common scenario for high-
speed signal recovery. This operating mode causes the Rx-PLL to lose lock if the data sampling
position is shifted too close to the center of a jitter distribution and thus, only allows for correct
BER measurements in the lower tail region of bathtub curves. However, this operating mode also
produces a random phase of the recovered clock as soon as the lock state is reached. Hence,
a DNL error of the BIJM system also affects the measurement result randomly. Over multiple
bathtub measurements it is thus observed as the random effect modeled in section 5.4.3. As test
channel, the 1m S-ATA crossover cable is again utilized. Figure 10.14 shows the statistical spread
of TJpp estimates over varying N with the Rx-PLL in lock-to-data mode (solid boxes). Using
equation (5.30) (uniform DJ) without the DNL error effect, the right dashed boxes are obtained.
Due to the present DNL error, especially the statistical spread is underestimated. This can be
solved by using equation (5.38) together with the tables of coefficients 5.6 and 6.3, and assuming
an additional DNL error of σDN L =0.1 UI. The obtained results are the dashed boxes in the center.
As a final analysis, in figure 10.15 the influence of varying channel length (RG-58 coaxial ca-
ble) on the measured TJ timing budget is shown. The solid lines correspond to median values of
estimated TJpp values with N =107 samples and the test pattern 08CEFhex . Black crosses mark
approximated true medians obtained from the N =1011 measurements, and dotted lines highlight
the expected worst case error under a uniform DJ assumption. As expected, the TJ budget gener-
ally increases with larger cable length and thus indicates the increase of ISI. For lock-to-reference
mode, the uniform DJ assumption is violated at large cable lengths due to additional noise influ-
153
10. A N FPGA BASED D IAGNOSTIC TOOL FOR J ITTER M EASUREMENT AND O PTIMIZATION
0.42 0.42
0.38 0.38
[UI]
[UI]
pp
pp
TJ
TJ
0.36 0.36
0.34 0.34
0.32 0.32
1M 10M 100M 1G 100G 1M 10M 100M 1G 100G
Sample Size N Sample Size N
F IGURE 10.14.: Estimated TJpp of a 1m S-ATA crossover cable with Rx-PLL in lock-to-data
mode (solid left boxes). Worst case error estimations assume uniform DJ us-
ing (5.30) (right dashed boxes) and (5.38) (center dashed boxes) which addi-
tionally includes the effect of process variations σDN L =0.1UI. ∆Pt =102 .
0.6
QN
0.55
Pat = 08CEFHEX
0.5
lock−to−data
0.45
TJpp
0.4
0.35
0.3
0.25
lock−to−reference
0.2 −1 0 1
10 10 10
Cable length [m]
F IGURE 10.15.: TJpp medians of tail fitted estimates (N =107 , circles), exact measurements
(N =1011 , crosses) and expected worst case uniform DJ (triangles) over varying
cable length. ∆Pt =103 .
ences. The same effect is also observed for very small cable lengths due to steep bathtub tails, as
discussed in the previous section with figure 10.7. In lock-to-data mode, overall TJ increases sig-
nificantly. The included Rx-PLL now superimposes its own jitter with unknown characteristic and
thus, violates the uniform DJ assumption as well. Also, measurement results suffer from larger
variations, as they are generally more affected by cable induced jitter.
10.4. Summary
An FPGA based diagnostic tool for total jitter estimation in high-speed interfaces has been devel-
oped. It allows for quantifying the timing budget caused by a system under test and thus, can be
used for testing the quality of transmission channels or optimizing the parameter configuration of
interface structures. Jitter measurements at 3Gb/s with a sample size of N =108 and R=128 delay
steps in a unit interval require approximately 2s of test time.
After an initial demonstration of the jitter measurement and sQN fitting principle in figure 10.5,
the TJpp values of different RG-58 cables were determined in figure 10.6 using the developed
diagnostic tool. The inherent jitter of the reference clock (TJpp ≈0.15 UI) was also quantified with
154
10.4. S UMMARY
loopback measurements and represented in figure 10.7. As an example for parameter optimization
with the presented diagnostic tool, the built-in Rx equalizers and Tx buffers of the FPGA were
configured together with a 5m RG-58 coaxial cable. This allowed to decrease the TJ timing
budget down to TJpp <0.3UI.
Further, the extrapolation error of sQN and QN fitting methods was investigated and compared
against theoretical worst case errors from sections 5.4 and 6.5. In this context the sinusoidal,
uniform and quadratic curve DJ shapes were experimentally confirmed to be well suited for pure
clock jitter, ISI, and ISI plus external noise affected channels (figures 10.11, 10.12 and 10.13).
When the receiver PLL was operated in normal lock-to-data mode, the additional effect of random
process variations could be made visible, which also allowed to apply the DNL error model in
figure 10.14. As a final analysis, in figure 10.15 the influence of varying RG-58 coaxial cable
length on the measured TJ timing budget was shown, together with the predicted worst case errors
under an ISI dominated DJ assumption. Results showed, that this condition is not fulfilled for large
cable lengths due to additional noise influences, as well as for very small cable lengths due to the
system limitations. Also in lock-to-data mode the overall TJ increases and changes the observed
DJ characteristic.
Parts of this chapter have also been published in [C7,C8].
155
11. Conclusion
This work concludes with an overall summary and overview to the key results achieved throughout
the thesis. A brief outlook to future directions is given at the end.
In this thesis, first the scaled Q-normalization (sQN) method for jitter and BER analysis was pre-
sented and realized in chapter 4. It is based on the Gaussian quantile function (equation (3.10)),
which linearizes the tails of jitter distributions and thus, allows for Gaussian tail fitting and extra-
polation. The Q-function was embedded into an efficient optimization scheme where a simple
recursion achieves a very fast exploration of the three-dimensional search space. With this re-
cursion, the sQN method automatically determines the best suited tail part for fitting. Thus, it
represents a clear improvement to other methods where the tail part is predefined in a conservative
manner or must be identified using an additional algorithm.
The extrapolation error of the basic sQN principle has been investigated in section 4.2, where
the major causes for a degraded performance have been identified as small sample sizes N , worst
case combinations of RJ and DJ, as well as Gaussian-like DJ shapes. However, the resulting
extrapolation error is always positive biased and thus pessimistic, which is a further beneficial
property of algorithms based on quantile normalization in general.
From the basic sQN principle, two optimized algorithmic scenarios have been derived which
allow for improved error behavior and robustness. The ĉ1.2 scenario in section 4.3 recommends
a minimum probability interval ∆Pt ≥102 for outermost tail selection. This parameter avoids
outliers caused by statistical tail variations. Further, the scenario combines fitness measures based
on both regression length and error to achieve an optimized error behavior. Another algorithmic
scenario, Qth,c , is based on a constant threshold Qmin in scaled Q-domain and thus, defines the
Gaussian tail region in terms of standard deviations beside the model mean. This representation
form allows for a flexible tail choice. As a trade-off, Qmin =−1.2 has been identified to achieve
acceptable accuracy with reduced risk for outlier occurrence.
Both algorithmic approaches improve the estimation performance compared to the basic sQN
principle. For example, with a typical uniform DJ, N =106 and worst case test distributions
(σRJ /ADJ =1/4), the error bias is still <2% with an overall error <3% in more than 97.5%
of the cases (confidence level a>0.95). The performance of the ĉ1.2 scenario is equivalent to Qth,c
for N ≥106 , and outperforms it at smaller sample sizes. This is due to a larger error variation of
Qth,c when a fit is performed at the outermost tail part.
In chapter 5, hardware design aspects for the sQN method were investigated. The basic idea was
to highlight properties of the proposed method when used together with test equipment, diagnostic
tools or built-in jitter measurement (BIJM) systems. Unlike simulations, these systems introduce
a discretization effect, which divides a distribution into R time intervals or bins per UI. As key
parameters for a system design, both the sample size N as well as the discrete number of bins
R were shown to cause fundamental limiting effects with respect to analyzable test distributions.
In order to correctly fit distribution tails, hence, equations were derived that specify minimum
requirements for tails.
157
11. C ONCLUSION
First the tail parameters of fitted test distributions were characterized in section 5.1 using the
polynomial equations (5.3) and (5.5). They allow for changing forth and back between the vari-
ables before (σRJ , ADJ ) and the obtained tail parameters after (At , σt , µt ) distribution synthesis.
The coefficients in table 5.1 and equation (5.3) specify tail amplitudes At obtained with the sQN
fitting method. The two parameters σt and µt describe minimum requirements for sQN tail fitting,
and are also valid for the conventional Q-normalization (QN) method described in chapter 6.
Requirements with respect to minimum tail amplitudes At,min were investigated in section 5.2.
For the ĉ1.2 algorithm a conservative threshold with N and the design parameter ∆Pt were given
in equation (5.10). For the Qth,c algorithm instead, equation (5.13) led to an exact solution, and
further highlighted the missing link to the ĉ1.2 algorithm based on ∆Pt .
Limitations introduced by the discrete time resolution with R number of bins per UI were in-
vestigated in section 5.3. This problem can also be represented in terms of a minimum analyzable
standard deviation σt,min of Gaussian tails. With equation (5.20) for ĉ1.2 and equation (5.23) for
the Qth,c algorithmic scenario, the minimum value σt,min is determined which can be fitted cor-
rectly by each of the algorithms. Selection charts that aid in identifying the design parameter ∆Pt
in figure 5.9 (ĉ1.2 ) or both Qmin and ∆Pt in figure 5.11 (Qth,c ) were provided as well. Note that
the parameters in both charts must also fulfill the previously mentioned requirements with respect
to the minimum tail amplitude (equation (5.10) or (5.13)), and outlier suppression with ∆Pt ≥102
chosen as large as possible.
In section 5.4 the combined influence of sample size and time discretization on the extrapolation
error of the sQN method was quantified. Therefore, the empirical relation (5.30) was derived
together with the tables of coefficients 5.4 (ĉ1.2 ) and 5.5 (Qth,c ). It describes bias, spread and
their combined influence on extrapolation error. The empirical relation is given in terms of a
two-dimensional function of sample size N and the variable product σRJ,min ·R. This product
is a measure for the bin density along a distribution tail. The empirical relation aids a designer
in finding an optimum performance trade-off between the required accuracy of a BIJM and the
hardware expense in terms of key parameters N and R. The obtained results clearly highlight a
better performance for the ĉ1.2 algorithm.
The additional effect of process variations in a jitter measurement system was investigated in
section 5.4.3. It is a typical result of timing mismatches inside the measurement system, and was
modeled as differential non-linearity (DNL) error with standard deviation σDN L . As a resulting
effect, the statistical spread of extrapolated tails increases significantly and thus, becomes the
major contributor to overall error. Thus, a well suited empirical equation (5.38) together with
included DNL error effect was derived, and according coefficients specified in table 5.6.
The effect of error oscillations at lowest bin densities was also investigated in section 5.4.2. If
a certain distribution shape is known, equation 5.35 allows to determine error maximums and to
adjust N and R accordingly. For the general case of unknown distribution shapes, σRJ,min ·R≥2
can avoid such oscillations.
Section 5.5 presented two typical design examples to highlight the calculation steps involved
with optimized BIJM designs. The first example assumed a jitter diagnosis scenario with a mini-
mum RJ tail of σt,min =0.01UI and a realizable number of R=128 bins per UI. First, the minimum
amplitude At,min was determined to correctly specify ∆Pt , which guaranteed for correct fitting
behavior. With equation (5.30) the sQN method achieved a worst case error bias Emed ≤2.0%
as well as an estimation loss EL ≤5.4% if the BIJM design was not affected by process varia-
tions. Otherwise, with σDN L =0.05UI and equation (5.38) the error increased to Emed ≤2.3% and
EL ≤7.4%.
The second design example focused on production testing with very stringent requirements on
the test time tt ≈20ms. With 14 parallel counters, the assumed system would be able to collect only
158
11.1. R ESULTS S UMMARY
N ≈26k jitter values within the given test time, but still achieve an extrapolation error EL <15%.
The performance of the proposed and previously published tail fitting principles based on Gaus-
sian quantile normalization was compared in chapter 6. These include the scaled Q-normalization
(sQN) method from chapter 4 and 5, the conventional Q-normalization method (QN), as well as
higher order polynomial methods (QP2, QP3 QP4) for tail fitting and extrapolation. The opti-
mization scheme in figure 6.1 was first derived to give a unifying, generalized perspective on
the compared methods. An efficient implementation of this scheme included the use of a fast
Levinson-Durbin recursion. Optimum parameter regions were derived using the ∆Pt parameter
for outer tail part selection and led to the configuration in table 6.1.
A comprehensive performance evaluation was first carried out for typical simulator environ-
ments by assuming R=Rsim =3.3·105 . With a varying sample size N and DJ type, figures 6.7-6.10
highlighted the different characteristics for each of the investigated methods. As a fundamental
result, the QN method showed the same beneficial property of a strictly positive error bias as the
sQN method. This is due to the asymptotic behavior of tails in Q-domain as already observed with
the sQN method in section 4.2.1. Hence, extrapolation results are always pessimistic. This is a
clear advantage compared to higher order polynomial methods (QP2, QP3, QP4), which achieved
acceptable accuracy only for certain DJ types or for a small range of distributional shapes. Third
and fourth order polynomials suffered from a large statistical variation of results, especially with
RJ dominant distributions where the Q-tails approximately follow a linear course. This problem
is generally due to an over-fitting effect.
A comprehensive comparison with the proposed sQN method (ĉ1.2 optimization based on ∆Pt )
was carried out in figures 6.11-6.15. The sQN method clearly achieved best performance as long
as the fitting condition for tail amplitudes in equation (5.10) was correctly met. Although the QN
method is less accurate, it offers the advantage of approximately 35 times faster tail fits, as was
also shown previously in figure 4.9. With this complementary property and the pessimistic tail
extrapolation, conventional QN is well suited for tail fitting and thus, also becomes a candidate for
hardware designs. The influence of such a limited time discretization RRsim on the extrapola-
tion error was investigated in figure 6.16. It basically showed a linear increase of the extrapolation
error over a large range of R for both QN and sQN methods.
In order to allow hardware designers to choose between the better suited algorithm alternative
for a jitter measurement system, the extrapolation performance of the QN method was evaluated in
section 6.5, equivalent to the sQN error analysis in section 5.4. Resulting coefficients were given
in table 6.2 for the empirical relation (6.6) without additional process variations, and in table 6.3
for the empirical relation (6.7) with included process variations as differential non-linearity error.
Hardware performance of the QN method was highlighted with the continued sQN design exam-
ples from section 5.5. Results showed that, although the QN method is generally less accurate
than sQN, it is also less affected by differential non-linearity error.
For a jitter measurement system using the QN method, the previously derived design equations
from chapter 5 can be applied again. Hence, equation (5.3) and condition (5.10) guarantee for a
minimum amplitude At,min , while equation (5.20) and the selection chart in figure 5.9 guarantee
for a minimum RJ standard deviation σt,min to be fitted correctly.
The flexible architecture of the sQN optimization scheme was highlighted in chapter 7, where a
generalized version let to a scheme for arbitrary non-Gaussian tail fitting. In this context, the gen-
eralized Gaussian distribution (GGD) function was introduced for tails with arbitrary exponential
power law behavior. A GGD uses an additional shape parameter α, and includes the Gaussian
distribution as a special case (α=2). It is thus fully consistent with the existing RJ-DJ model.
For simulations where Rsim is sufficiently large, the GGD method achieved acceptable accu-
racy, although it suffered from a slightly negative or optimistic error bias. The reason is a general
overestimation of shape parameters, as was shown in figure 7.6. For hardware scenarios with lim-
159
11. C ONCLUSION
ited R, additional error further degraded the performance and let to unreliable extrapolations. This
use case is thus not recommended. A performance comparison in figure 7.9 clearly highlighted
the advantage of the generalized method when used together with heavy tailed test distributions
(α≤2), instead of the sQN method from chapter 4. Since the proposed method was implemented
for use with a broad range of test distributions, further improvements may be achieved if spe-
cific shapes or test distributions are known and the fitting algorithm only focuses on these special
characteristics.
Chapter 8 provided a first application example for the sQN method when used with system
behavioral simulations. Therefore, an accurate CPLL model was implemented as an enhanced
version of a prior event-driven approach. It allowed for dynamic run-time variations of parameters
and included a gain regulator pole as well as a VCO noise model. On an Intel Core Duo 2.2GHz
laptop, the model was able to gather up to 20k jitter values per second simulation time.
Initial simulations in figure 8.4 compared the closed loop phase noise PSD with measurement
data at different parameter configurations, and showed an excellent agreement. Only spectral spurs
at multiples of 50MHz, could not be reflected as they were not modeled. Two different simula-
tion methods, one based on the sQN method and the other on a spectral method, determined jitter
transfer functions of the modeled PLL. Resulting curves were again compared with measurements
in figures 8.7-8.9. Both methods correctly identified the measured cut-off frequency of transfer
functions. However, the spectral method better reflected the peaking behavior in the cut-off re-
gion, while the sQN method was instead able to correctly track curves obtained with different test
patterns. This is due to the same analysis principle as used in measurements.
A second application example for the sQN method was given in chapter 9, together with an
adaptive algorithm for jitter tolerance analysis of high-speed PLLs. It consists of an adaptive
recursion (equation (9.1)) as well as a mechanism for automatic sample size adaptation (figure 9.2),
to efficiently determine jitter tolerance curves.
In the example, the algorithm was applied to the PLL model from chapter 8. After deriving a
well suited set of parameters (table 9.2), the performance of four different algorithmic combina-
tions was investigated. These included use of the QN and sQN fitting methods as well as their
realizations without sample size adaptation (c-QN and c-sQN).
Results highlighted a general decrease of the total number of required jitter samples N by a
factor of 2-3, if the automatic sample size adaptation (QN, sQN) is utilized. The smallest com-
putational effort was achieved with QN and c-QN. The smallest number of iterations instead, was
given without sample size adaptation (c-QN and c-sQN), because their constant error bias does
not influence the adaptation process. Hardware simulations with a reduced number of bins R (ta-
ble 9.4 and figure 9.6) indicated that the developed algorithm can also be applied to hardware jitter
measurements. Finally, in figure 9.7 simulated jitter tolerance curves were also compared against
measurements from the according test structure and highlighted an excellent matching.
The third application example for tail fitting methods was given in chapter 10, where an FPGA
based diagnostic tool for TJ estimation in high-speed interfaces was implemented. The mea-
surement routine used an efficient algorithm, which realizes jitter measurements at 3Gb/s with a
sample size of N =108 and R=128 in approximately 2s of test time. Using this diagnostic tool,
the TJpp values of different RG-58 cables were determined in figure 10.6. The inherent jitter of
the reference clock was quantified with loopback measurements and indicated TJpp ≈0.15 UI. As
an example for parameter optimization, the built-in Rx equalizers and Tx buffers of the FPGA
were configured in figures 10.8-10.10 together with a 5m RG-58 coaxial cable. For the best case
with an included DFE, the TJ timing budget was decreased over 28% below TJpp <0.3UI.
The extrapolation error of sQN and QN fitting methods was investigated and compared against
theoretical worst case errors from sections 5.4 and 6.5 respectively. This allowed to experimentally
160
11.2. O UTLOOK
confirm the sinusoidal, uniform and quadratic curve DJ shapes as well suited for clock jitter, ISI,
and ISI plus additional noise affected channels (figures 10.11, 10.12 and 10.13). The DNL error
model was also successfully applied in figure 10.14 with jitter measurements from the Rx-PLL
operating in lock-to-data mode.
11.2. Outlook
The sQN method from chapter 4 was shown to achieve an excellent accuracy of extrapolated
tails, combined with a fast and flexible tail fitting procedure. The derived equations and empirical
relations are primarily intended for use with a broad variety of test distributions and DJ types. For
further improvements with respect to certain distribution shapes, it would thus be interesting to
focus on specific test cases, where one of the tail parameters is for example known or can easily
be approximated.
The residual analysis of fitted quantiles in section 4.1.3 highlighted a non-constant and corre-
lated error structure for the outermost tail region in Q-domain. This makes the linear regression
of quantiles still sub-optimal with respect to a maximum likelihood tail search [20]. However,
an additional weighting and de-correlation of errors requires an excessive computational demand,
and is thus not feasible. Nevertheless it would be interesting to compare the sQN error with the
theoretical performance maximum for tail fitting methods. Such comparisons might for example
become possible with fitting methods based on maximum likelihood approaches.
The generalized fitting principle in chapter 7 suffers from accuracy with a reduced number
of bins R or a large shape parameter α. A detailed analysis of different fitness measures and
their combinations, as well as the focus on a few specific tail shapes could possibly improve the
performance.
Finally, with the given empirical error analysis of the presented sQN and QN methods, the
extrapolation accuracy in all kind of future BIJM designs, diagnostic tools or measurement devices
can be predicted as long as they provide jitter distributions. Thus, the derived empirical relations
may also find a direct application in preliminary concept studies of novel systems.
161
A. Figure Data Files
The following lists denotes the MATLAB scripts which were used to post-process the C simula-
tions, in order to generate the figures and tables of coefficients throughout this work. The scripts
can be found in the sub-folders specified in the last column. All simulation data and results are
documented on an appended DVD, available upon request at:
[email protected]
163
A. F IGURE DATA F ILES
164
Table Section Page MATLAB File Folder
5.1 5.1.1 62 app_tj_analysis 3/6/
5.2 5.1.1 64 app_tj_analysis 3/6/
5.4 76
5.4.1 stim_ber_N_RJmin_looped 4/3/
5.5 77
5.6 5.4.3 80 stim_ber_N_RJmin_DNL_looped 4/3/
6.2 6.5 105 stim_ber_N_RJmin_looped 4/3/
6.3 6.5 106 stim_ber_N_RJmin_DNL_looped 4/3/
TABLE A.3.: List of System-C testbenches for simulations and MATLAB post-processing files.
165
Own Publications
[C1] S. Erb and W. Pribyl, “An Accurate and Efficient Method for BER Analysis in High-
Speed Communication Systems", IEEE European Conf. on Circuit Theory and Design
(ECCTD’09), Aug. 2009.
[C2] S. Erb and W. Pribyl, “A Behavioral Modeling Approach for Jitter Analysis in Charge-Pump
PLLs", Austrian Workshop on Microelectronics (AUSTROCHIP’09), Oct. 2009.
[C3] S. Erb and W. Pribyl, “Comparison of Jitter Decomposition Methods for BER Analysis of
High-Speed Serial Links", IEEE Symp. on the Design and Diagnostics of Electronic Circuits
and Systems (DDECS’10), Apr. 2010.
[C4] S. Erb and W. Pribyl, “Design and Performance Considerations for an On-Chip Jitter Anal-
ysis System", IEEE Int. Symp. on Circuits and Systems (ISCAS’10), May 2010.
[C5] S. Erb and W. Pribyl, “An Approach to Generalized Jitter and BER Analysis", Austrian
Workshop on Microelectronics (AUSTROCHIP’10), Oct. 2010.
[C6] S. Erb and W. Pribyl, “A Method for Fast Jitter Tolerance Analysis of High-Speed PLLs",
IEEE Conf. Design Automation and Test in Europe (DATE’11), Mar. 2011.
[C7] S. Erb, M. Stadler and W. Pribyl, “An FPGA based Diagnostic Tool for Jitter Optimization in
Serial High-Speed Transceivers", IEEE Ph.D. Research in Microelectronics & Electronics
(PRIME’11), Jul. 2011, submitted for publication in Feb. 2011.
[C8] S. Erb and W. Pribyl, “Design Specification for BER Analysis Methods using Built-in Jitter
Measurements", IEEE Trans. VLSI Systems, submitted for publication in Oct. 2010.
[C9] S. Erb, “Method and Device for Predicting a Figure of Merit from a Distribution", U.S.
Patent Application, US2010/0 246 650 A1, Sep. 30, 2010.
167
Bibliography
[1] P. Acco, M. P. Kennedy, C. Mira, B. Morley, and B. Frigyik, “Behavioral Modeling of
Charge Pump Phase Locked Loops,” in IEEE Int. Symp. Circuits and Systems (ISCAS’99),
1999, pp. 375–378.
[3] Agilent Tech. (2003, Feb.) Jitter Analysis Techniques for High Data Rates. White Paper.
Cited 2010-09-17. [Online]. Available: www.agilent.com
[4] S. Ahmed and T. Kwasniewski, “Efficient Simulation of Jitter Tolerance for All-Digital
Data Recovery Circuits,” IEEE Midwest Symp. Circuits and Systems (MWSCAS’07), pp.
1070–1073, Aug. 2007.
[5] S. Ali, “Basics of Chip-to-Chip and Backplane Signaling,” IEEE Solid-State Circuits Conf.
Tutorial (ISSCC’08), Feb. 2008.
[7] R. E. Best, Phase-Locked Loops - Design, Simulation and Applications, 5th ed. New York
(NY): McGraw-Hill, 2003.
[8] W. Beyene, C. Madden, J.-H. Chun, H. Lee, Y. Frans et al., “Advanced Modeling and Ac-
curate Characterization of a 16 Gb/s Memory Interface,” IEEE Trans. Advanced Packaging,
vol. 32, no. 2, pp. 306–327, May 2009.
[9] L. Bizjak, “Development of a PLL Blocks Library for accurate Time-Domain simulations
and Clock Analysis Software,” Master’s thesis, Università degli studi di Udine, I, 2005.
[10] G. E. P. Box and M. E. Muller, “A Note on the Generation of Random Normal Deviates,”
Ann. Math. Statistics, no. 29, pp. 610–611, 1958.
[11] M. Brownlee, “Low Noise Clocking for High Speed Serial Links,” Ph.D. dissertation, Ore-
gon State University, US, 2006.
[13] Y. Cai, S. Werner, G. Zhang, M. Olsen, and R. Brink, “Jitter Testing for Multi-Gigabit
Backplane SerDes - Techniques to Decompose and Combine Various Types of Jitter,” IEEE
Int. Test Conf. (ITC’02), pp. 700–709, Oct. 2002.
[14] Y. Cai, B. Laquai, and K. Luehman, “Jitter Testing for Gigabit Serial Communication Trans-
ceivers,” IEEE Design & Test of Computers, vol. 19, no. 1, pp. 66–74, Jan.-Feb. 2002.
169
B IBLIOGRAPHY
[15] B. Casper and F. O’Mahony, “Clocking Analysis, Implementation and Measurement Tech-
niques for High-Speed Data Links - A Tutorial,” IEEE Trans. Circuits Syst. I, vol. 56, no. 1,
pp. 17–39, Jan. 2009.
[17] A.-S. Chao and S.-J. Chang, “A Jitter Characterizing BIST with Pulse-Amplifying Tech-
nique,” IEEE Asian Test Symp. (ATS’09), pp. 379–384, Nov. 2009.
[18] K.-H. Cheng, J.-C. Liu, C.-Y. Chang, S.-Y. Jiang, and K.-W. Hong, “Built-in Jitter Mea-
surement Circuit With Calibration Techniques for a 3-GHz Clock Generator,” IEEE Trans.
VLSI Systems, vol. PP, no. 99, pp. 1–11, Jun. 2010.
[20] M. J. Crawley, The R Book. Chichester (GB): John Wiley & Sons, 2007.
[21] B. Daniels, R. Farrell, and G. Baldwin, “Arbitrary Order Charge Approximation Event
Driven Phase Lock Loop Model,” IET Irish Signals and Systems Conf. (ISSC), Jul. 2004.
[22] N. Da Dalt, M. Harteneck, C. Sandner, and A. Wiesbauer, “Numerical Modeling of PLL Jit-
ter and the Impact of its Non-White Spectrum on the SNR of Sampled Signals,” Southwest
Symp. Mixed-Signal Design (SSMSD’01), pp. 38–44, Feb. 2001.
[23] N. Da Dalt and C. Sandner, “Introduction to PLL Jitter Definitions,” presentation slides,
Feb. 2001.
[24] N. Da Dalt, “Cheetah CDR L90 v2 Measurements,” Infineon Technologies, Jan. 2007, con-
fidential.
[25] N. Da Dalt, “Theory and Implementation of Digital Bang-Bang Frequency Synthesizers for
High Speed Serial Data Communications,” Dissertation, Technische Hochschule Aachen,
D, 2007.
[26] A. R. DiDonato and A. H. Morris, Jr., “Computation of the Incomplete Gamma Function
Ratios and their Inverse,” ACM Trans. Math. Softw., vol. 12, no. 4, pp. 377–393, 1986.
[27] Q. Dou and J. Abraham, “Jitter Decomposition by Time Lag Correlation,” IEEE Int. Symp.
Quality Electronic Design (ISQED’06), Mar. 2006.
[29] J. Eckle-Kohler and M. Kohler, Eine Einführung in die Statistik und ihre Anwendungen.
Berlin (D): Springer, 2008.
[30] W. F. Egan, Phase-Lock Basics, 2nd ed. Hoboken (NJ): John Wiley & Sons, 2007.
[31] M. Evans, N. Hastings, and B. Peacock, Statistical Distributions, 2nd ed. Hoboken (NJ):
John Wiley & Sons, 1993.
[32] Y. Fan, Y. Cai, L. Fang, A. Verma, W. Burchanowski et al., “An Accelerated Jitter Tolerance
Test Technique on ATE for 1.5GB/s and 3GB/s Serial-ATA,” IEEE Int. Test Conf. (ITC’06),
pp. 1–10, Oct. 2006.
170
B IBLIOGRAPHY
[33] Y. Fan and Z. Zilic, “Accelerating Jitter Tolerance Qualification for High Speed Serial In-
terfaces,” IEEE Int. Symp. Quality of Electronic Design (ISQED’09), pp. 360–365, Mar.
2009.
[34] A. Fog. (2008, Feb.) “C++ Library of Pseudo Random Number Generators”. Cited
2008-10-09. [Online]. Available: www.agner.org/random/
[35] A. Frisch, “Jitter Measurement System and Method,” U.S. Patent 2002/0 106 014 A1, Aug.
8, 2002.
[36] F. M. Gardner, Phase Lock Techniques, 3rd ed. Hoboken (NJ): John Wiley & Sons, 2004.
[37] T. Georges, “Non-Gaussian timing jitter statistics of controlled solitons,” in Optical Fiber
Communications (OFC’96), Feb. 1996, pp. 232–233.
[38] F. Ghenassia, Ed., Transaction Level Modeling with SystemC. Dordrecht (NL): Springer,
2005.
[39] G. H. Golub and C. F. Van Loan, Matrix Computations, 3rd ed. Baltimore (MD): Johns
Hopkins University Press, 1996.
[40] V. Grigoryan, C. Menyuk, and R.-M. Mu, “Calculation of Timing and Amplitude Jitter
in Dispersion-Managed Optical Fiber Communications using Linearization,” J. Lightwave
Technology, vol. 17, no. 8, pp. 1347–1356, Aug. 1999.
[41] T. Grötker, S. Liao, G. Martin, and S. Swan, System Design with SystemC. Boston (MA):
Kluwer Academic Publishers, 2002.
[42] A. Hajimiri and T. Lee, The Design of Low Noise Oscillators. Dordrecht (NL): Kluwer
Academic Publishers, 1999.
[43] B. Ham, “Methodologies for Jitter and Signal Quality Specification,” INCITS, Tech. Rep.,
Jun. 2005.
[44] Y. Hamaizi and A. El-Akrmi, “Soliton Propagation in Fiber Systems,” in ICTON Mediter-
ranean Winter Conf. (ICTON-MW’09), Dec. 2009, pp. 1–4.
[45] P. K. Hanumolu, “Design Techniques for Clocking High Performance Signaling Systems,”
Ph.D. dissertation, Oregon State University, US, 2006.
[47] P. Hanumolu, B. Casper, R. Mooney, G.-Y. Wei, and U.-K. Moon, “Jitter in High-Speed
Serial and Parallel Links,” IEEE Int. Symp. Circuits and Systems (ISCAS’04), vol. 4, pp.
IV–425–428, May 2004.
[49] S. Haykin, Communication Systems, 4th ed. Hoboken (NJ): John Wiley & Sons, 2001.
171
B IBLIOGRAPHY
[52] D. Hong, C.-K. Ong, and K.-T. Cheng, “Bit-Error-Rate Estimation for High-Speed Serial
Links,” IEEE Trans. Circuits Syst. I, vol. 53, no. 12, pp. 2616–2627, Dec. 2006.
[53] D. Hong, “Efficient Test Methodologies for High-Speed Serial Links,” Ph.D. dissertation,
University of California, Santa Barbara, US, 2008.
[54] D. Hong and K.-T. Cheng, “An Accurate Jitter Estimation Technique for Efficient High
Speed I/O Testing,” IEEE Asian Test Symp. (ATS’07), pp. 224–229, Oct. 2007.
[55] D. Hong and K.-T. Cheng, “Bit-Error Rate Estimation for Bang-Bang Clock and Data Re-
covery Circuit in High-Speed Serial Links,” IEEE VLSI Test Symp. (VTS’08), pp. 17–22,
May 2008.
[56] D. Hong, C.-K. Ong, and K.-T. Cheng, “BER Estimation for Serial Links Based on Jitter
Spectrum and Clock Recovery Characteristics,” IEEE Int. Test Conf. (ITC’04), pp. 1138–
1147, Oct. 2004.
[57] J.-C. Hsu and C. Su, “BIST for Measuring Clock Jitter of Charge-Pump Phase-Locked
Loops,” IEEE Trans. Instrumentation and Measurement, vol. 57, no. 2, pp. 276–285, Feb.
2008.
[58] J.-L. Huang, “A Random Jitter Extraction Technique in the Presence of Sinusoidal Jitter,”
IEEE Asian Test Symp. (ATS’06), pp. 318–326, Nov. 2006.
[59] M. Hueda, D. Crivelli, H. Carrer, and O. Agazzi, “Parametric Estimation of IM/DD Opti-
cal Channels Using New Closed-Form Approximations of the Signal PDF,” J. Lightwave
Technology, vol. 25, no. 3, pp. 957–975, Mar. 2007.
[60] K. Ichiyama, M. Ishida, T. J. Yamaguchi, and M. Soma, “Novel CMOS Circuits to Measure
Data-Dependent Jitter, Random Jitter, and Sinusoidal Jitter in Real Time,” IEEE Trans.
Microw. Theory Tech., vol. 56, no. 5, pp. 1278–1285, May 2008.
[62] Nitrophy Analog Core Reference Manual, Infineon Technologies, 2007, confidential.
[63] A. Jantsch, Modeling Embedded Systems and SoC’s. US: Elsevier Science, 2003.
[65] K. Jenkins, A. Jose, Z. Xu, and K. Shepard, “On-chip Circuit for Measuring Jitter and
Skew with Picosecond Resolution,” in Proc. of IEEE Int. Conf. Integrated Circuit Design
and Technology (ICICDT’08), Jun. 2008, pp. 257–260.
172
B IBLIOGRAPHY
[66] S.-Y. Jiang, K.-H. Cheng, and P.-Y. Jian, “A 2.5-GHz Built-in Jitter Measurement System
in a Serial-Link Transceiver,” IEEE Trans. VLSI Systems, vol. 17, no. 12, pp. 1698 –1708,
Dec. 2009.
[67] S. G. Johnson. (2009, Nov.) “The NLopt Nonlinear-Optimization Package”. MIT. Cited
2010-01-31. [Online]. Available: ab-initio.mit.edu/nlopt
[68] J. Kim, “On-Chip Measurement of Jitter Transfer and Supply Sensitivity of PLL/DLLs,”
IEEE Trans. Circuits and Systems II: Express Briefs, vol. 56, no. 6, pp. 449–453, Jun. 2009.
[69] K. K. Kim, J. Huang, Y.-B. Kim, and F. Lombardi, “Analysis and Simulation of Jitter
Sequences for Testing Serial Data Channels,” IEEE Trans. Industrial Informatics, vol. 4,
no. 2, pp. 134–143, May 2008.
[70] J. M. Kizer and C. J. Madden, “Method for Estimating RJ and DJ,” U.S. Patent US
2006/0 059 392 A1, Mar. 16, 2006.
[71] R. Koenker, Quantile Regression. New York (NY): Cambridge University Press, 2005.
[72] M. Kossel and M. Schmatz, “Jitter Measurements of High-Speed Serial Links,” IEEE De-
sign & Test of Computers, vol. 21, no. 6, pp. 536–543, Nov.-Dec. 2004.
[73] M. Kubicek, “In-System Jitter Measurement Using FPGA,” in 20th Int. Conf. Radioelek-
tronika, Apr. 2010, pp. 1–4.
[74] K. Kundert. (2006, Aug.) Modeling Jitter in PLL-based Frequency Synthesizers. Cited
2008-05-27. [Online]. Available: www.designers-guide.org
[75] A. Kuo, T. Farahmand, N. Ou, S. Tabatabaei, and A. Ivanov, “Jitter Models and Measure-
ment Methods for High-Speed Serial Interconnects,” IEEE Int. Test Conf. (ITC’04), pp.
1295–1302, Oct. 2004.
[76] A. Kuo, R. Rosales, T. Farahmand, S. Tabatabaei, and A. Ivanov, “Crosstalk Bounded Un-
correlated Jitter (BUJ) for High-Speed Interconnects,” IEEE Trans. Instrumentation and
Measurement, vol. 54, no. 5, pp. 1800–1810, Oct. 2005.
[77] P. Larsson, “A Simulator Core for Charge-Pump PLLs,” IEEE Trans. Circuits Syst. II,
vol. 45, no. 9, pp. 1323–1326, Sep. 1998.
[78] T. Lee and A. Hajimiri, “Oscillator Phase Noise: A Tutorial,” IEEE J. Solid-State Circuits,
vol. 35, no. 3, pp. 326–336, Mar. 2000.
[79] Y. Lee, C.-Y. Yang, N.-C. Cheng, and J.-J. Chen, “An Embedded Wide-Range and High-
Resolution Clock Jitter Measurement circuit,” in IEEE Conf. Design, Automation and Test
in Europe (DATE’10), 2010, pp. 1637–1640.
[80] H. Le Gall, “Estimating of the Jitter of a Clock Signal,” U.S. Patent 7 487 055, Feb. 3, 2009.
[81] H. Le Gall, “Jitter Estimation Circuit,” ST Microelectronics, Mar. 2011, private communi-
cation, confidential.
[82] M. P. Li, Jitter, Noise, and Signal Integrity at High-Speed. Boston (MA): Prentice Hall,
2007.
[83] M. Li, “Jitter Challenges and Reduction Techniques at 10 Gb/s and Beyond,” IEEE Trans.
Advanced Packaging, vol. 32, no. 2, pp. 290–297, May 2009.
173
B IBLIOGRAPHY
[84] M. Li, J. Wilstrup, R. Jessen, and D. Petrich, “A New Method for Jitter Decomposition
Through its Distribution Tail Fitting,” IEEE Int. Test Conf. (ITC’99), pp. 788–794, Sep.
1999.
[85] M. Li, J. Wilstrup, R. Jessen, and D. Petrich, “Method and Apparatus for Analyzing Mea-
surement,” U.S. Patent 6 298 315, Oct. 2, 2001.
[86] T. Mak, M. Tripp, and A. Meixner, “Testing Gbps Interfaces without a Gigahertz Tester,”
IEEE Design & Test of Computers, vol. 21, no. 4, pp. 278–286, July-Aug. 2004.
[87] M. Mansuri, A. Hadiashar, and C.-K. K. Yang, “Methodology for On-Chip Adaptive Jitter
Minimization in Phase-Locked Loops,” IEEE Trans. Circuits and Systems II: Analog and
Digital Signal Processing, vol. 50, no. 11, pp. 870–878, Nov. 2003.
[88] MathWorks Inc. (2009, Mar.) “Optimization Toolbox User’s Guide V4.2 (R2009a)”.
MATLAB Documentation. [Online]. Available: www.mathworks.com
[89] “Physical Layer Performance: Testing the Bit Error Ratio (BER),” Technical Article,
Maxim Inc., Sep. 2004.
[90] S. McClure, “Digital Jitter Measurement and Separation,” Master’s thesis, Texas Tech Uni-
versity, US, 2006.
[91] J. A. McNeill and D. Ricketts, The Designer’s Guide to Jitter in Ring Oscillators. New
York (NY): Springer, 2010.
[92] S. E. Meninger, “Low Phase Noise, High Bandwidth Frequency Synthesis Techniques,”
Ph.D. dissertation, Massachusetts Institute of Technology, US, 2007.
[93] M. Miller, “Estimating Total Jitter Concerning Precision, Accuracy and Robustness,” De-
signCon, Feb. 2007.
[94] M. Miller, “Measuring Components of Jitter,” U.S. Patent 7 516 030, Apr. 7, 2009.
[95] M. Miller. (2007) Normalized Q-scale analysis: Theory and background. EDN. Cited
2011-04-12. [Online]. Available: www.edn.com
[96] M. Müller, R. W. Stephens, and R. McHugh, “Total Jitter Measurement at Low Probability
Levels, Using Optimized BERT Scan Methods,” White Paper, Agilent Technologies, 2005.
[97] P. Muller and Y. Leblebici, “Jitter Tolerance Analysis of Clock and Data Recovery Circuits
using Matlab and VHDL-AMS,” in Proc. of Forum on Design Languages (FDL’05), 2005.
[98] F. Nan, Y. Wang, F. Li, W. Yang, and X. Ma, “A Better Method than Tail-fitting Algorithm
for Jitter Separation Based on Gaussian Mixture Model,” J. of Electronic Testing: Theory
and Applications, vol. 25, no. 6, pp. 337–342, Dec. 2009.
[99] R. Nonis, “Phase Noise Modeling in PLL Frequency Synthesizers,” Master’s thesis, Uni-
versità degli studi di Udine, I, 2002.
[100] K. Nose, M. Kajita, and M. Mizuno, “A 1-ps Resolution Jitter-Measurement Macro Using
Interpolated Jitter Oversampling,” IEEE J. Solid-State Circuits, vol. 41, no. 12, pp. 2911–
2920, Dec. 2006.
174
B IBLIOGRAPHY
[101] C.-K. Ong, D. Hong, K.-T. Cheng, and L.-C. Wang, “Jitter Spectral Extraction for Multi-
Gigahertz Signal,” IEEE Asia - South Pacific Design Automation Conf. (ASP-DAC’04), pp.
298–303, Jan. 2004.
[102] C.-K. Ong, D. Hong, K.-T. Cheng, and L.-C. Wang, “A Clock-Less Jitter Spectral Analysis
Technique,” IEEE Trans. Circuits Syst. I, vol. 55, no. 8, pp. 2263–2272, Sept. 2008.
[103] A. V. Oppenheim and R. W. Schafer, Discrete-Time Signal Processing, 3rd ed. New York
(NY): Pearson, 2010.
[104] N. Ou, T. Farahmand, A. Kuo, S. Tabatabaei, and A. Ivanov, “Jitter Models for the Design
and Test of Gbps-Speed Serial Interconnects,” IEEE Design & Test of Computers, vol. 21,
no. 4, pp. 302–313, July-Aug. 2004.
[105] H. Pang, J. Zhu, and W. Huang, “Jitter Decomposition by Fast Fourier Transform and Time
Lag Correlation,” in IEEE Int. Conf. Communications, Circuits and Systems ICCCAS, Jul.
2009, pp. 365–368.
[107] M. Perrott, “Fast and Accurate Behavioral Simulation of Fractional-N Frequency Synthe-
sizers and Other PLL/DLL Circuits,” in IEEE Design Automation Conf. (DAC’02), Jun.
2002, pp. 498–503.
[108] M. Perrott, M. Trott, and C. Sodini, “A Modeling Approach for Σ-∆ Fractional-N Fre-
quency Synthesizers Allowing Straightforward Noise Analysis,” IEEE J. Solid-State Cir-
cuits, vol. 37, no. 8, pp. 1028–1038, Aug. 2002.
[109] M. H. Perrott. (2008, Apr.) “CppSim System Simulator Package V4”. MIT. Cited
2009-06-09. [Online]. Available: www.cppsim.com
[110] M. H. Perrott, “Digital Phase-Locked Loops,” IEEE Solid-State Circuits Conf. Tutorial
(ISSCC’08), Feb. 2008.
[111] A. Popovici, “Fast Measurement of Bit Error Rate in Digital Links,” IEE Proc. Communi-
cations, Radar and Signal Processing, vol. 134, no. 5, pp. 439–447, Aug. 1987.
[112] M. J. Porsani and T. J. Ulrych, “Levinson-Type Algorithms for Polynomial Fitting and
for Cholesky and Q-Factors of Hankel and Vandermonde Matrices,” IEEE Trans. Signal
Processing, vol. 43, no. 1, pp. 63–70, Jan. 1995.
[113] M. J. D. Powell, “The BOBYQA Algorithm for Bound Constrained Optimization with-
out Derivatives,” Department of Applied Mathematics and Theoretical Physics, Cambridge
England, Tech. Rep., 2009.
[114] J. Proakis, Digital Communications, 4th ed. New York (NY): McGraw-Hill, 2001.
[115] R.-D. Reiss and M. Thomas, Statistical Analysis of Extreme Values, 2nd ed. Basel (CH):
Birkhäuser, 2001.
[116] F. Scholz. (2008, May) Applications of the Noncentral t-Distribution. Stat 498B Industrial
Statistics. Cited 2011-04-12. [Online]. Available: www.stat.washington.edu/fritz/
[117] F. Scholz, “Nonparametric Tail Extrapolation,” Boeing Information & Support Services,
ISSTECH-95-014, 1995. [Online]. Available: www.stat.washington.edu/fritz/
175
B IBLIOGRAPHY
[118] Serial ATA Specification, Serial ATA International Organization Std., Rev. 2.6, 2006.
[119] M. Shimanouchi, “An Approach to Consistent Jitter Modeling for Various Jitter Aspects
and Measurement Methods,” IEEE Int. Test Conf. (ITC’01), pp. 848–857, Oct.-Nov. 2001.
[120] M. Shimanouchi, M. Li, and D. Chow, “New Modeling Methods for Bounded Gaussian
Jitter (BGJ)/Noise (BGN) and their Applications in Jitter/Noise Estimation/Testing,” IEEE
Int. Test Conf. (ITC’09), pp. 1–8, 2009.
[121] K. Shu and E. Sánchez-Sinencio, CMOS PLL Synthesizers. Boston (MA): Springer Sci-
ence, 2005.
[123] R. W. Stephens, “Jitter Analysis: The dual-Dirac model, RJ/DJ, and Q-Scale,” White Paper,
Agilent Technologies, Dec. 2004.
[124] R. Stephens, “Separation of Random and Deterministic Components of Jitter,” U.S. Patent
7 149 638, Dec. 12, 2006.
[127] V. Stojanovic, “Channel Limited High-Speed Serial links - Modeling, Analysis and De-
sign,” Ph.D. dissertation, Stanford University, US, 2004.
[128] J. Sun, M. Li, and J. Wilstrup, “A Demonstration of Deterministic jitter (DJ) Deconvo-
lution,” IEEE Instrumentation and Measurement Technology Conf. (IMTC’02), vol. 1, pp.
293–298, May 2002.
[129] S. Sunter and A. Roy, “On-chip Digital Jitter Measurement, from Megahertz to Gigahertz,”
IEEE Design & Test of Computers, vol. 21, no. 4, pp. 314–321, July-Aug. 2004.
[130] S. Tabatabaei and A. Ivanov, “Embedded Timing Analysis: A SoC Infrastructure,” IEEE
Design & Test of Computers, vol. 19, no. 3, pp. 22–34, May-June 2002.
[131] C.-C. Tsai and C.-L. Lee, “An On-Chip Jitter Measurement Circuit for the PLL,” IEEE
Asian Test Symp. (ATS’03), pp. 332–335, Nov. 2003.
[132] S. Vamvakos, C. Werner, and B. Nikolic, “Phase-Locked Loop Architecture for Adaptive
Jitter Optimization,” in Int. Symp. Circuits and Systems (ISCAS ’04), vol. 4, May 2004, pp.
IV–161–164.
[133] R. C. Walker, “Designing Bang-Bang PLLs for Clock and Data Recovery in Serial Data
Transmission Systems,” in Phase-Locking in High-Performance Systems: From Devices to
Architectures, B. Razavi, Ed. IEEE Press, 2003, pp. 34–45.
[134] Z. Wang, “An Analysis of Charge-Pump Phase-Locked Loops,” IEEE Trans. Circuits Syst.
I, vol. 52, no. 10, pp. 2128–2138, Oct. 2005.
176
B IBLIOGRAPHY
[137] Xilinx Inc. (2008, Nov.) Virtex-5 FPGA RocketIO GTX Transceiver User’s Guide.
ug198.pdf. Cited 2011-02-15. [Online]. Available: www.xilinx.com
[138] Xilinx Inc. (2009, Oct.) ML50x Evaluation Platform User’s Guide. ug347.pdf. Cited
2011-02-15. [Online]. Available: www.xilinx.com
[139] T. Yamaguchi, H. Hou, K. Takayama, D. Armstrong, M. Ishida et al., “An FFT-based Jitter
Separation Method for High-Frequency Jitter Testing with a 10x Reduction in Test Time,”
IEEE Int. Test Conf. (ITC’07), pp. 1–8, Oct. 2007.
[140] T. Yamaguchi, M. Soma, M. Ishida, H. Musha, and L. Malarsie, “A New Method for Testing
Jitter Tolerance of SerDes Devices Using Sinusoidal Jitter,” IEEE Int. Test Conf. (ITC’02),
pp. 717–725, Oct. 2002.
[141] J. Yin and L. guang Zeng, “A Statistical Jitter Tolerance Estimation Applied for Clock and
Data Recovery Using Oversampling,” IEEE Region 10 Conf. (TENCON’06), pp. 1–4, Nov.
2006.
[142] I. Zamek and S. Zamek, “Definitions of Jitter Measurement Terms and Relationships,” IEEE
Int. Test Conf. (ITC’05), Nov. 2005.
[143] J. Zhu and W. Huang, “Jitter Analysis and Decomposition Based on EMD/HHT in
High-Speed Serial Communications,” in Proc. of IEEE Int. Conf. Testing and Diagnosis
(ICTD’09), Apr. 2009, pp. 1–4.
177